01.12.2012 Views

Photodriven heterogeneous charge transfer with transition-metal ...

Photodriven heterogeneous charge transfer with transition-metal ...

Photodriven heterogeneous charge transfer with transition-metal ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

This article was published as part of the<br />

2009 Renewable Energy issue<br />

Reviewing the latest developments in renewable<br />

energy research<br />

Guest Editors Professor Daniel Nocera and Professor Dirk Guldi<br />

Please take a look at the issue 1 table of contents to access<br />

the other reviews.


CRITICAL REVIEW www.rsc.org/csr | Chemical Society Reviews<br />

<strong>Photodriven</strong> <strong>heterogeneous</strong> <strong>charge</strong> <strong>transfer</strong> <strong>with</strong> <strong>transition</strong>-<strong>metal</strong><br />

compounds anchored to TiO 2 semiconductor surfacesw<br />

Shane Ardo and Gerald J. Meyer*<br />

Received 24th October 2008<br />

First published as an Advance Article on the web 1st December 2008<br />

DOI: 10.1039/b804321n<br />

A critical review of light-driven interfacial <strong>charge</strong>-<strong>transfer</strong> reactions of <strong>transition</strong>-<strong>metal</strong><br />

compounds anchored to mesoporous, nanocrystalline TiO 2 (anatase) thin films is described. The<br />

review highlights molecular insights into <strong>metal</strong>-to-ligand <strong>charge</strong> <strong>transfer</strong> (MLCT) excited states,<br />

mechanisms of interfacial <strong>charge</strong> separation, inter- and intra-molecular electron <strong>transfer</strong>, and<br />

interfacial <strong>charge</strong>-recombination processes that have been garnered through various spectroscopic<br />

and electrochemical techniques. The relevance of these processes to optimization of solar-energyconversion<br />

efficiencies is discussed (483 references).<br />

1. Introduction<br />

A Rationale<br />

Hoffert has elegantly documented recent power needs on the<br />

terawatt (TW = 10 12 W) scale. 1,2 As the worldwide rate of energy<br />

expenditure is directly related to the number of people on Earth,<br />

the population growth experienced over the last quarter-century<br />

is staggering: a 45% increase which equates to roughly two<br />

billion people and 6 TW of additional power (B63% increase). 3<br />

This coupled <strong>with</strong> the urbanisation of third-world and industrialized<br />

nations and cities has led to an increase in the demand for<br />

fuel that has subsequently driven gas and oil prices to record<br />

highs. 3 Regardless of their price, the continued use of fossil fuels<br />

cannot be a long-term solution as they come from a limited stock<br />

and the deleterious environmental consequences of their combustion<br />

have become self-evident. Thus, the numbers alone, i.e.<br />

Johns Hopkins University, 3400 North Charles Street, Baltimore, MD<br />

21218, USA<br />

w Part of the renewable energy theme issue.<br />

Shane Ardo was born in San<br />

Francisco, California in 1977.<br />

He obtained a BS in mathematics<br />

from Towson University<br />

in 1999 and an MS in<br />

nutrition and food science<br />

from UM-College Park in<br />

2005. Since joining Johns<br />

Hopkins University for a<br />

Shane Ardo<br />

PhD program in chemistry,<br />

Shane has been awarded an<br />

MS degree and was recently<br />

selected to describe renewable<br />

energy sources at the inaugural<br />

Eaton E. Lattman lecture<br />

series. He currently studies<br />

molecular, photo-induced processes at nanocrystalline TiO2<br />

interfaces for application in dye-sensitized solar cells and photoelectrosynthetic<br />

hydrogen formation. Shane also enjoys soccer,<br />

hiking, and camping <strong>with</strong> his fiance´e and friends.<br />

population, energy demand, and fuel prices, do not convey the<br />

severity of the problem. Concern should be elicited as the ice-core<br />

data over the past three-quarters-of-a-million years correlates<br />

temperature <strong>with</strong> greenhouse gas concentration 5,6 and current<br />

atmospheric CO2 levels of 4380 ppm 4 exceed any values attained<br />

over this same time period. 5 Further, outside of natural photosynthesis,<br />

there exist no means by which our society could<br />

significantly lower the concentration of CO2. The increased<br />

average global temperature and rates of glacial melting measured<br />

over the last few decades are telling signs. 7 There is real reason for<br />

concern. Regardless of one’s opinion on the causes of global<br />

climate change, it is very difficult to argue <strong>with</strong> two key points:<br />

(1) humans need to conserve energy and (2) commercially viable<br />

and sustainable energy conversion processes need to be discovered,<br />

designed, and developed.<br />

The motivation for this review stems from the urgent need<br />

for inexpensive and sustainable materials that can be used for<br />

solar energy conversion. Hoffert and co-workers concluded<br />

that in order to avoid catastrophic planetary changes Earth<br />

will require at least 10 terawatts of carbon-neutral energy by<br />

Gerald (Jerry) J. Meyer was<br />

born in Oconomowoc Wisconsin<br />

in 1962. He received a BS in<br />

chemistry and mathematics from<br />

SUNY-Albany and a PhD in<br />

chemistry from UW-Madison.<br />

After a post-doctoral appointment<br />

at UNC-Chapel Hill, he<br />

joined the faculty at Johns<br />

Hopkins University in 1991<br />

where he is now the<br />

Bernard N. Baker Professor of<br />

Chemistry <strong>with</strong> a joint appointment<br />

in the Materials Science &<br />

Jerry Meyer Engineering Department. In<br />

addition to his interests in environmental<br />

chemistry and solar energy conversion, Jerry enjoys long<br />

distance running, tennis, cooking, gardening, hiking, and spending<br />

time <strong>with</strong> his wife, Lisa, and daughters, Caroline and Jillian.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 115


Fig. 1 A schematic depicting a champion dye-sensitized solar cell (DSSC) illustrating the approximate relative energetics of individual electron<strong>transfer</strong><br />

reactions along <strong>with</strong> their corresponding rate constants or current densities. The steps highlighted in this review are shown as blue Roman<br />

numerals and subcategorized by capitalized letters: (I) sensitizer light absorption; (II) excited-state electron injection; (III) regeneration of the<br />

oxidized sensitizer by an electron donor in the electrolyte; (IV) <strong>charge</strong> recombination of TiO 2 electrons, TiO 2(e )s, to (A) oxidized sensitizers or<br />

(B) oxidized donors. Adapted from Fig. 9 of ref. 11.<br />

the year 2050, which was approximately equal to the worldwide<br />

energy requirement in the year 1998. 1,2 They also described<br />

the pitfalls of a ‘wait-and-see’ approach and<br />

recommended immediate action. It has now been dubbed the<br />

Terawatt Challenge. 8 The sun is the one source that on its own<br />

could supply the world’s projected energy demand and in a<br />

sustainable fashion. 4 To put it in perspective, the amount of<br />

solar energy reaching the earth in one day could power the<br />

planet for an entire year. 8,9 Remaining is the challenge of<br />

harvesting and storing this energy in a cost-effective way. It is<br />

our assumption that molecular approaches to this challenge<br />

will ultimately be most successful. The relative ease by which the<br />

spectroscopic and electrochemical properties of molecules can be<br />

tuned through synthetic manipulation allows for many minute<br />

variations on solar-energy-conversion schemes to be rapidly<br />

studied. Additionally, chemical bonds afford large energy storage<br />

capacities, i.e. energy densities, and power densities that exceed<br />

those obtainable from most other storage methods.<br />

B Background<br />

When we started our research program at Johns Hopkins<br />

University in 1991, molecular approaches to solar-energy conversion<br />

were solely of academic interest. The hard fact was that<br />

the most efficient ‘‘molecular solar cells’’ were comprised of cold<br />

water running over illuminated black paint. In the same year<br />

much progress in the field was realized when Grätzel and<br />

O’Regan reported an order-of-magnitude increase in solar<br />

light-to-electrical power conversion efficiency from dye-sensitized<br />

solar cells (DSSCs). 10 Their significant advance was to replace<br />

the planar electrode materials of the past <strong>with</strong> high surface area,<br />

mesoporous, nanocrystalline semiconductor thin films. The<br />

actual surface area for sensitizerz binding was up to three<br />

z Since a photocurrent is generated <strong>with</strong> light of lower energy than the<br />

bandgap of TiO 2, the chromophoric dyes are referred to as sensitizers,<br />

a term that we will use throughout this review.<br />

orders-of-magnitude larger than the geometric surface area,<br />

which is critical for solar harvesting <strong>with</strong> molecular compounds.<br />

11 Today, confirmed efficiencies in excess of 10% have<br />

been established. 12<br />

The general mechanisms for light-to-electrical power conversion<br />

in DSSCs were developed in early sensitization studies<br />

of planar and single-crystal semiconductors. Mechanisms like<br />

that shown in Fig. 1 can be found in many excellent reviews on<br />

this area. 11,13 In short: (I) light is absorbed by a sensitizer to<br />

form a molecular excited state; (II) the excited state may inject<br />

an electron into the semiconductor thus causing <strong>charge</strong> separation;<br />

(III) the oxidized sensitizer is ‘‘regenerated’’ by an<br />

external electron donor. Once the electron has performed<br />

useful work in the external circuit, it returns to a counter<br />

electrode where it reduces the oxidized electron donor. Hence<br />

the solar cell is termed ‘regenerative’ as all oxidation chemistry<br />

at the dye-sensitized electrode is reversed at a dark counter<br />

electrode such that no net chemistry occurs. It is now possible<br />

to include rate constants and/or current densities for many of<br />

these processes as well as for (IV) the unwanted <strong>charge</strong><br />

recombination of TiO 2 electrons to: (A) oxidized sensitizers;<br />

or (B) oxidized donors in the electrolyte. There are a tremendous<br />

number of details in an operational Gra¨tzel-type cell and<br />

the values in Fig. 1 represent a good starting point for their<br />

general comparison. However, the time scales and current<br />

densities are often misleading as they may be specific to a<br />

certain class of sensitizers or electrolytes and/or may be<br />

abstracted from experimental data obtained in the absence<br />

of some components of the operational Gra¨tzel-type cell.y<br />

y The term Grätzel-type cell was chosen so as to highlight the distinction<br />

between the DSSCs employed in the seminal Nature paper—the use of<br />

mesoporous, nanocrystalline TiO2 (anatase) thin-film electrodes—<br />

and those in previous DSSC fabrications. Although historically accurate,<br />

this distinction will only be noted in the Introduction section and not<br />

throughout this review as preferred by Professor Gra¨tzel. 14<br />

116 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


In this review we provide some of the details that have<br />

arisen from recent research of <strong>heterogeneous</strong>, <strong>charge</strong>-<strong>transfer</strong><br />

processes involved in the transduction of energy in TiO 2-based<br />

DSSCs. They predominantly deal <strong>with</strong> actions occurring at<br />

mesoporous, nanocrystalline TiO 2 (anatase) electrodes sensitized<br />

to visible light <strong>with</strong> <strong>transition</strong>-<strong>metal</strong> coordination compounds.<br />

Space limitations prevented us from including results<br />

obtained in the active areas of research <strong>with</strong> organic and<br />

quantum dot sensitizers. The review highlights molecular insights<br />

into the interfacial, <strong>charge</strong>-<strong>transfer</strong> processes I–IV<br />

that have been garnered through various spectroscopic and<br />

electrochemical measurements. As this review illustrates, many<br />

of the details remain poorly understood. Not<strong>with</strong>standing,<br />

numerous interesting and informative studies have been<br />

performed in order to probe electrolyte- or counter electrodebased<br />

phenomena 15 as well as <strong>charge</strong> transport through<br />

mesoporous films. 16–20 These will be discussed only as they<br />

are relevant to processes I–IV.<br />

Understanding the operation of a Gra¨tzel cell is not the only<br />

reason to study excited states and interfacial electron <strong>transfer</strong> at<br />

semiconductor interfaces on the molecular level, a point often<br />

missed by reviewers in this area. Indeed the spirit of using<br />

inexpensive processing and non-toxic, abundant, high surfacearea<br />

materials for solar-energy conversion is exactly on track. 13 It<br />

is a sound approach toward practical solutions to the Terawatt<br />

Challenge that could ultimately provide future generations the<br />

relatively low-cost power that we enjoy today, but in a sustainable<br />

fashion. 8 To build on the success of the Gra¨tzel cell and<br />

develop low-cost materials capable of solar-energy conversion<br />

and storage is just one of many motivations for understanding<br />

interfacial <strong>charge</strong> <strong>transfer</strong> in precise molecular detail.<br />

2. Solar harvesting <strong>with</strong> <strong>metal</strong>-polypyridyl<br />

compounds<br />

One sun of solar irradiance at an Air Mass of 1.5 (AM1.5) and<br />

under standard, U.S. atmospheric conditions (1000 W m 2 )is<br />

often taken as an average irradiance and spectral distribution of<br />

sunlight in the United States. The spectrum is available in<br />

downloadable format form the National Renewable Energy<br />

Laboratory (NREL) website. 21 The fraction of light that is<br />

absorbed by a DSSC is wavelength dependent and is often called<br />

the light harvesting efficiency (LHE) or the more generally<br />

preferred IUPAC name, absorptance (a(l)). 22 The absorptance<br />

of a monolayer of sensitizers anchored to a flat surface is related<br />

to (a) the molar extinction coefficient of the sensitizer, e(l), and<br />

(b) the surface area occupied by the dye, ADye, i.e. the footprint:<br />

aðlÞ ¼1<br />

IðlÞ<br />

IoðlÞ ¼ 1 10 AbsorbanceðlÞ ; where ð1Þ<br />

Absorbance ðlÞ ¼log IoðlÞ<br />

IðlÞ<br />

¼ 1000 eðlÞ<br />

1<br />

ADye<br />

where Io(l) is the intensity of the incoming incident light and I(l)<br />

is the intensity of the light transmitted through the sample.<br />

Calculations show that even monolayers of phthalocyanines<br />

and porphyrins, which have among the highest extinction coefficients<br />

known, absorb far less than 1% of the 1 sun, AM1.5<br />

spectrum on planar surfaces. 23 This underscores the need for<br />

ð2Þ<br />

high surface-area materials to increase the LHE of a molecular<br />

monolayer of sensitizers.<br />

The effectiveness of a solar cell is measured by its power<br />

output. This value is the product of its current and voltage. Thus,<br />

determination of the solar cell’s current–voltage relationship<br />

often aids in assessing its performance, Fig. 2. The lightto-electrical<br />

power conversion efficiency of a solar cell (Z) isthe<br />

product of the open-circuit photovoltage (Voc), short-circuit<br />

photocurrent (isc), and fill factor (FF) divided by the product of<br />

the incident irradiance (Po)andtheareaofthesolarcell(Acell). 24<br />

Z ¼ iscVocFF<br />

PoAcell<br />

Very often P o is set to be 1 sun of AM1.5 solar irradiation,<br />

i.e. 1000 W m 2 . V oc is the maximum Gibbs free energy that one<br />

can abstract from a regenerative solar cell, while i sc is the<br />

maximum rate that <strong>charge</strong> can flow through the external circuit.<br />

The long-wavelength absorption edge sets a thermodynamic limit<br />

to the Voc and the LHE can be used to calculate the maximum<br />

possible isc; thus, assuming FF = 1, the largest possible Z can<br />

be estimated all from a simple absorption measurement of the<br />

solar cell. The optimal isc assumption is, <strong>with</strong>in experimental<br />

uncertainty, realized in champion DSSCs. 25 However, the<br />

spectroscopically estimated maximum V oc values are much<br />

greater than those that have been observed experimentally.<br />

A subtlety is that the i sc of a solar cell is directly related to its<br />

absorptance (a), but not its absorbance. The absorbances and<br />

absorptances are approximately equal at low sensitizer concentrations<br />

but differ significantly at the high sensitizer concentrations<br />

used in champion DSSCs. Therefore, the normalized<br />

photocurrent action spectrum, i.e. a plot of the incident<br />

photon-to-current efficiency (IPCE), or external quantum efficiency,<br />

as a function of excitation wavelength, should coincide<br />

<strong>with</strong> the normalized sensitizer absorptance spectrum. One is<br />

Fig. 2 Typical current–voltage curve for a champion DSSC under<br />

approximately 1 sun, AM1.5 illumination (0.998 suns). Labeled are the<br />

short-circuit photocurrent (isc), open-circuit photovoltage (Voc), and<br />

power point (PP) along <strong>with</strong> its corresponding photovoltage (V PP)<br />

and photocurrent (iPP). The fill factor (FF) is the area of the shaded<br />

region, which is bounded by the V PP and i PP, divided by the area of the<br />

region outlined by the dashed line, which is bounded by the Voc and isc.<br />

The curve in magenta represents the power as a function of voltage in<br />

arbitrary units further illustrating that the PP coincides <strong>with</strong> the<br />

condition of maximum power output. Adapted from Fig. 6 of ref. 26.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 117<br />

ð3Þ


often interested in not only the monochromatic absorptance but<br />

the integrated, a(l)-weighted solar flux divided by the total 1<br />

sun, AM1.5 photon flux as well. The latter represents the<br />

overall percentage of solar light absorbed where the numerator<br />

serves as an upper limit to the i sc of the DSSC.<br />

The FF can be related to i sc and V oc through the corresponding<br />

values at the power point (PP):<br />

FF ¼ iPPVPP<br />

ð4Þ<br />

iscVoc<br />

where the PP occurs at the maximum product of the cell<br />

output photovoltage and photocurrent obtained along the<br />

current–voltage curve, Fig. 2. While FF = 1 is ideal, such a<br />

value cannot be achieved due to various loss mechanisms such<br />

as <strong>charge</strong> recombination.<br />

A Orbitals and electronic <strong>transition</strong>s<br />

The <strong>metal</strong>-to-ligand <strong>charge</strong> <strong>transfer</strong> (MLCT) excited states of dp 6<br />

coordination compounds have emerged as the most efficient for<br />

solar harvesting and sensitization of wide-bandgap semiconductor<br />

materials. As the name implies, light absorption promotes an<br />

electron from the Metal d orbitals to the Ligand p* orbitals,<br />

d(p) - p*. 27–29 A number of electric-dipole-allowed Charge-<br />

Transfer <strong>transition</strong>s are observed which give rise to intense<br />

absorption bands in the visible region <strong>with</strong> moderate extinction<br />

coefficients. There is no formal spin for each excited state due to<br />

heavy-atom spin–orbit coupling from the <strong>transition</strong>-<strong>metal</strong><br />

center (especially for 4d and 5d <strong>metal</strong>s). 30,31 Crosby et al. have<br />

proposed that the excited state is accurately described by solely<br />

the symmetry label of the molecular point group to which it<br />

belongs, corresponding to an irreducible representation, and not<br />

the spin and an orbital individually. 30 The effects of spin–orbit<br />

coupling must be introduced in order to rationalize the<br />

relative oscillator strengths and absorption spectra of<br />

M(bpy) 3 2+ (M = Fe II ,Ru II and Os II ) compounds, where bpy<br />

is 2,2 0 -bipyridine.<br />

The classical example of a compound <strong>with</strong> such <strong>transition</strong>s is<br />

Ru(bpy)3 2+ which is arguably the most well-studied, coordination<br />

compound. Its lowest-energy state is three-fold symmetric<br />

and is best described by the symmetry label D3,Fig.3.Basedon<br />

the Franck–Condon principle, immediately following excitation<br />

the initial excited state ought to possess the same structural<br />

symmetry as the ground state. 32–34 Thus, in the absence of<br />

Jahn–Teller distortions or solvent-induced fluctuations, the<br />

initial, Franck–Condon excited state formed via an MLCT<br />

<strong>transition</strong> in Ru(bpy)3 2+ could consist of a delocalized electronic<br />

wavefunction on all three bpy ligands each formally possessing<br />

1/3 of an electronic <strong>charge</strong>. Monitoring the conversion of<br />

this excited-state from D3 to C2 symmetry is a non-trivial task,<br />

although some evidence supports the notion that conversion<br />

occurs by T2 dephasing. 35,36 Based on the absence of an electric<br />

dipole for D3 symmetry molecules and minor, but clearly<br />

observable, solvent-dependent, ground-state MLCT absorption<br />

features, the initial excited-state electron is thought to localize<br />

on a single bpy. 37 Time-resolved resonance Raman spectroscopy<br />

of Ru(bpy) 3 2+ shows clear evidence for localization on<br />

nanosecond and longer time scales. 38 This localized excited state<br />

has the reduced-symmetry designation C2 and an estimated<br />

dipole moment of B10 Debye. 37,39<br />

Demas and colleagues have shown that intersystem crossing<br />

to a manifold of relaxed, MLCT excited states occurs <strong>with</strong> a<br />

quantum yield near unity in fluid solution, Fig. 4(a). 41–43<br />

Although not formally triplet or singlet in nature, the predominantly<br />

triplet character of the lowest-energy excited state,<br />

1E 0 , 44,45 and singlet character of the initial Franck–Condon<br />

state rationalizes why the <strong>transition</strong> between them is often<br />

termed intersystem crossing. It is for this reason that these states<br />

will be labeled as 3 MLCT and 1 MLCT, respectively, throughout<br />

this review. Crosby, Hager, and colleagues have shown that<br />

photoluminescence (PL) arises from three closely spaced electronic<br />

states. 46–50 Rapid thermal equilibrium between this manifold<br />

of states, okT in energy apart, happens such that PL occurs<br />

from what appears to be a single thermally equilibrated excited<br />

state, or thexi state. 51,52 Yersin et al. discovered evidence for two<br />

more highest-energy states by temperature-dependent emission<br />

polarization experiments and labeled them per the D 3 0 double<br />

symmetry group, which takes into account the spin–orbit<br />

Fig. 4(b). 44,45 These <strong>transition</strong>s are generally supported by those<br />

Fig. 3 Molecular-orbital diagrams for Ru(L)6 2+ -type compounds in their ground state <strong>with</strong>: GS-Oh) octahedral, Oh, symmetry; or GS-D3)<br />

reduced D 3 symmetry, like for Ru(bpy) 3 2+ . Also shown are excited-state molecular-orbital diagrams for: 3 MLCT-D3) the initial, Franck–Condon<br />

excited state formed under the ground-state D3 symmetry, where the excited electron is delocalized equally over each ligand; and 3 MLCT-C2) the<br />

excited state possessing reduced C 2 symmetry where the excited electron is localized on one ligand. Taken from Fig. 2 of ref. 40.<br />

118 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 4 (A) A Jablonski-type energy diagram for Ru(bpy)3 2+ illustrating its manifold of thermally equilibrated excited states, i.e. the thexi state.<br />

The quantum yield for intersystem crossing, f ISC, is approximately unity. Taken from Scheme 1 of ref. 54. (B) The relative energy levels for the<br />

excited states of Ru(bpy)3 2+ under the D3 0 double group, which takes spin–orbit coupling into consideration. Taken from Fig. 3 of ref. 44.<br />

obtained from computational Density Functional Theory (DFT)<br />

calculations. 53<br />

As many of the sensitizers employed in champion DSSCs are<br />

of the form cis-Ru(LL) 2X 2, where LL is a bpy-like ligand and X<br />

is a non-chromophoric ligand, their spectral differences and<br />

similarities to Ru(bpy) 3 2+ are discussed. When LL = bpy and<br />

X=NC the compound’s spectrum is solvatochromic and the<br />

Ru III/II reduction potential, E o (Ru III/II ), is more negative as<br />

compared to Ru(bpy)3 2+ . 55 This coupled <strong>with</strong> the relatively<br />

insensitive energetics of the p* orbitals of the bpy ligand leads<br />

to red-shifted absorption and emission maxima. These lowestenergy,<br />

actinic <strong>transition</strong>s are MLCTinnatureandresultinan<br />

electron on the bpy-based chromophoric ligand and a hole that is<br />

partially delocalized on the cyano ligands, thus greatly decreasing<br />

the Lewis basicity of the cyano ligands. The most efficient<br />

sensitizerforDSSCsiscalledN3,whereLL=4,4 0 -dicarboxylic<br />

acid-bpy (dcb) and X = SCN . 25 Although less solvatochromic<br />

than the cyano derivative, its visible absorption spectrum, and<br />

that of its ‘LL = bpy’ derivative, exhibit two well-resolved<br />

bands. It has been postulated that this occurs due to a shift in<br />

the electron density of the highest occupied molecular orbital<br />

(HOMO) from the Ru II -<strong>metal</strong> center to the isothiocyanate<br />

ligands. 56–58 By DFT it was calculated that B75% of the<br />

HOMO density resides on the isothiocyanate ligands and that<br />

B75% of this density resides on the sulfur atom.<br />

B Tuning of the absorption spectrum and redox properties<br />

An important aspect of dp 6 coordination compounds is that<br />

their colors can be widely tuned using synthetic chemistry. The<br />

MLCT absorption bands can be tuned in energy by altering<br />

the substituents on the bpy ligands or by controlling the extent<br />

of d(p)-p* back-bonding donation to nonchromophoric<br />

ligands. How these changes affect the photophysical properties<br />

of the compounds have been the subject of many investigations<br />

affording further insights into the factors that govern<br />

radiative and nonradiative excited-state decay. As just<br />

mentioned, the compound that has emerged as the most<br />

efficient sensitizer for DSSC application is N3. 25 N3 gains<br />

red absorption over Ru(dcb)3 2+ however at the expense of the<br />

E o (Ru III/II ). Although not generally vital to DSSCs, this loss in<br />

driving force for regeneration of the oxidized sensitizer would<br />

further limit the sensitizer’s ability to perform a ‘holy grail’ of<br />

chemistry: water oxidation. 59,60 However, in terms of DSSC<br />

light-to-electrical power conversion efficiency, N3 and closely<br />

related analogues remain unsurpassed, Fig. 5.z A similarly<br />

successful Ru II -based sensitizer, which is based on terpyridine<br />

rather than bpy, is the so called ‘black dye’: [Ru(tct)(NCS)3] ,<br />

where tct is 4 0 ,4 00 ,4 000 -tricarboxylic acid-tpy and tpy is<br />

2,2 0 :6 0 ,2 00 -terpyridine. It extends the spectral sensitivity of the<br />

solar cell significantly towards the red as compared to N3. 68<br />

However, a lower extinction coefficient throughout the visible<br />

region results in an overall less-efficient DSSC.<br />

The reduction potentials of the thexi state of the sensitizers,<br />

E o (Ru III/II* ) and E o (Ru II*/+ ), can be estimated using thermochemical<br />

cycles. 69,70 In many cases the spectroscopic and<br />

electrochemical data needed for such calculations can be<br />

measured in situ, i.e. for the sensitizer anchored to the semiconductor<br />

film. Previous studies have shown that molecules<br />

anchored to mesoporous, nanocrystalline TiO2, ZrO2, or<br />

Al2O3 thin films can be reversibly oxidized in standard electrochemical<br />

cells provided that the surface coverage exceeds a<br />

percolation threshold. 71–73 Cyclic voltammetry and spectroelectrochemistry<br />

are thus powerful in situ tools for determining<br />

formal reduction potentials and absorption spectra of relevant<br />

redox states. The excited-state reduction potential for<br />

the oxidation of the thermally equilibrated excited state,<br />

E o (Ru III/II* ), is calculated by the following equation:<br />

E o (Ru III/II* )=E o (Ru III/II ) DG ES (5)<br />

z Analogues where one or more of the carboxylic acid groups have<br />

been deprotonated, e.g. the dianion salt of N3 <strong>with</strong> tetrabutylammonium<br />

counterions (TBA + )—N719, 61 or one of the dcb ligands has<br />

been replaced by a more hydrophobic bipyridine ligand, e.g.<br />

Z907—<strong>with</strong> replacement by 4,4 0 -dinonyl-bpy 62–64 —and K19—<strong>with</strong><br />

replacement by 4,4 0 -bis(p-hexyloxystyryl)-bpy. 65–67<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 119


Fig. 5 The chemical structures of the most successful Ru II -based sensitizers employed in champion DSSCs.<br />

where DG ES is the free energy stored in the thexi state. 69,70,74,75<br />

This energy can be estimated by the PL onset or through a<br />

Franck–Condon lineshape analysis of the corrected PL spectrum.<br />

The reduction potential of the initially formed,<br />

Franck–Condon excited state can be calculated by substituting<br />

the excitation energy for DGES.<br />

As previously mentioned, Ru(bpy)3 2+ and most other trisheteroleptic<br />

Ru II compounds have redox and optical properties<br />

that are fairly insensitive to their environments. 37,76 However,<br />

this is not the case for ammine and cyano compounds of the<br />

type [M(bpy 0 )(X) 4] 2 ,2+ or [cis-M(bpy 0 ) 2(X) 2] 0,2+ ,M=Fe,<br />

Ru, or Os and X = CN or NH 3. 76 Outer-sphere interactions<br />

<strong>with</strong> the cyano ligands have a profound influence on E o (M III/II )<br />

and hence the color of the compound. [Ru(dcb)(CN)4] 2<br />

is<br />

highly solvatochromic; 78 the maximum of the lower-energy<br />

MLCT band of Ru(dcb)(CN)4/TiO2 was observed at 450<br />

10 nm in tetrahydrofuran and at 500 20 nm in dimethylformamide.<br />

78 The color change was due to a shift of E o (Ru III/II )<br />

<strong>with</strong> solvent. The complex maintained this solvatochromism<br />

upon attachment to mesoporous, nanocrystalline TiO2<br />

(anatase) thin films although the magnitude of the effect<br />

decreased. Solvent tuning altered the spectral responses of<br />

DSSCs based on these materials in a predictable way. For<br />

[Fe(bpy)(CN) 4] 2<br />

compounds, the excited-state reorganization<br />

energy in acetonitrile was found to be significantly larger on<br />

TiO2 than in fluid solution (l = 0.32 eV versus 0.10 eV,<br />

respectively). 77 This increased reorganization energy may<br />

be due to the restricted translational mobility of the semiconductor-bound<br />

iron compounds and the ambidentate<br />

Fe II –CN–Ti IV linkages. Interestingly, a recent Raman study<br />

has shown that when anchored to TiO2, the solvent reorganization<br />

energy of N3 decreased by a factor of six. 79 Further studies<br />

are needed to provide fundamental information on the solvation<br />

environment of similar semiconductor-bound molecules.<br />

A shortcoming of actinic sensitization by MLCT <strong>transition</strong>s<br />

is their relatively low extinction coefficients as compared to<br />

p - p* <strong>transition</strong>s often found in organic sensitizers. Thus<br />

6–10 mm thick films of nanocrystalline TiO2 are required for<br />

efficient solar harvesting and increased LHE <strong>with</strong> Ru II -based<br />

coordination compounds. This precludes the use of many<br />

classes of semiconductor materials that have inherently low<br />

surface areas. Ru(bpy)3 2+ has a molar extinction coefficient of<br />

about 15 000 M 1 cm 1 for its MLCT-based electronic <strong>transition</strong>s.<br />

80 In contrast, natural and synthetic organic pigments<br />

also absorb solar photons but <strong>with</strong> extinction coefficients that<br />

are often in excess of 200 000 M 1 cm 1 . 23 It has long been<br />

known that addition of substituents to bpy <strong>with</strong> low lying<br />

p orbitals (such as aromatics, esters, carboxylic acids, or<br />

unsaturated organics) can enhance MLCT extinction coefficients<br />

relative to unsubstituted bpy. 65–67,81–85 Interestingly,<br />

4 and 4 0 disubstitution of bpy has been found to increase<br />

these extinction coefficients more effectively than does disubstitution<br />

in the 5 and 5 0 positions. 86 The preparation of high<br />

extinction coefficient, heteroleptic N3 derivatives, where one<br />

of the dcb ligands is replaced by a 4,4 0 -disubstituted bpy is an<br />

extremely active area of research. 65–67,82,84,85,87<br />

We recently found that employing bpy ligands bridged in the<br />

3and3 0 positions by dithiolene is a viable alternative to the<br />

more traditional and widely pursued approach of introducing<br />

conjugated groups in the 4 and 4 0 positions. 81 Substituent<br />

effects in this position are not as well documented as they<br />

sterically force the two pyridyl rings out of planarity, behavior<br />

that can decrease the stability of the compound. This issue is<br />

circumvented <strong>with</strong> bridging ligands but at the expense of<br />

opening up the N–Ru–N bite angle thereby stabilizing<br />

ligand-field states and decreasing the excited-state lifetime.<br />

Nevertheless, it was notable that these first-derivative, MLCTdithiolene<br />

compounds have extinction coefficients for their lowest-energy<br />

<strong>transition</strong>s that are comparable to the highest ever<br />

reported based on Ru II (4,4 0 -disubstituted-bpy) compounds,<br />

4.4 10 4 M 1 cm 1 . In a similar absorption region the largest<br />

value for the dyes often employed in champion DSSCs, Fig. 5, is<br />

1.8 10 4 M 1 cm 1 for K19 66 and to the best of our knowledge<br />

no compounds exceed 3.9 10 4 M 1 cm 1 beyond 450 nm. 83,85<br />

Part of this success was that the dithiolene-bpy ligands<br />

have intraligand absorption bands, in addition to the MLCT<br />

absorption bands, in the visible region.<br />

An alternative strategy for increasing the LHE is to use<br />

nature’s antenna effect, Fig. 6. 88–94 Multiple pigments that are<br />

suitably arranged can absorb light and vectorally <strong>transfer</strong> their<br />

energy to a central pigment that can then inject an electron<br />

into the semiconductor. If the additional pigments do not<br />

increase the footprint of the sensitizer on the semiconductor<br />

surface, this is a method for enhancing the LHE. Indeed, the<br />

trinuclear Ru II sensitizer utilized in the celebrated 1991 Nature<br />

paper had been previously designed in Italy to function as an<br />

antennae. 91 An issue <strong>with</strong> the Ru(dcb)2(CN)2 group used as<br />

the energy <strong>transfer</strong> acceptor and surface anchor was the cis<br />

120 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 6 A scheme depicting an array of sensitizers bound to a planar<br />

TiO2 surface consisting of cis- and trans-[(Ru(bpy)2(pz))4(ina)] 8+ on<br />

the left and right, respectively, where pz is ambidentate pyrazine and<br />

ina is isonicotinic acid. The trans orientation may allow for increased<br />

absorptance, a, <strong>with</strong>out increasing the projected footprint of the<br />

sensitizer. Taken from Fig. 1 of ref. 95.<br />

geometry of the ambidentate cyano ligands, which resulted in<br />

a larger footprint as the number of Ru II pigments was<br />

increased. In this regard, a trans geometry is more preferred. 95<br />

The synthesis of molecules that function as antennae and their<br />

use in DSSCs continues to be an active area of research that<br />

may one day enable the efficient sensitization of planar<br />

semiconductor materials. 93<br />

C Excited-state time scales<br />

The excited-state lifetime of [Ru III (bpy)2(bpy )] 2+ * is B1<br />

microsecond in water. 96 The radiative rate constant is typically<br />

about two orders-of-magnitude smaller than the non-radiative<br />

rate constant and hence the excited-state lifetime is controlled<br />

by the latter. 96 Ru II - and Os II -polypyridyl excited states have<br />

been shown to follow Jortner’s Energy Gap Law, where the<br />

non-radiative rate constant increases exponentially <strong>with</strong> decreasing<br />

energy gap. 97–101 For this reason, it has proven to be<br />

difficult to prepare compounds that emit in the infrared region<br />

and have long-lived excited states. A large ligand-field splitting<br />

parameter is required for the observation of long lifetimes in<br />

this class of excited states. The presence of low-lying, ligandfield<br />

states can rapidly deactivate MLCT excited states and<br />

decrease excited-state lifetimes. A classical example of this is<br />

Fe(bpy)3 2+ which, until recently, was thought to be completely<br />

non-emissive due to rapid and quantitative internal conversion/intersystem<br />

crossing through ligand-field states.<br />

As described further below, one fascinating aspect of DSSCs<br />

is the ultrafast excited-state injection into the semiconductor<br />

which has been observed under many experimental conditions.<br />

102–119 It is therefore useful to describe the time scales<br />

on which Ru II -based coordination compounds undergo equilibration<br />

to their MLCT thexi states. Using transient absorption<br />

anisotropy measurements on Ru(bpy)3 2+ in acetonitrile,<br />

McCusker and colleagues have identified <strong>charge</strong>-localizing<br />

decoherence of the initial, Franck–Condon, D3-symmetrical<br />

excited state occurring <strong>with</strong> a lifetime of 59 fs. 36 The kinetics<br />

were proposed to be coupled to inertial solvent dynamics as<br />

the lifetimes were solvent dependent in nitrile solvents and<br />

ranged from 59 to 173 fs in an order expected based on such a<br />

hypothesis. The contradictory conclusion that formation of<br />

such a C2-symmetrical excited state occurs immediately upon<br />

light excitation can be disregarded assuming a decoherent<br />

mechanism for the randomization of the initially formed,<br />

D 3-symmetrical excited state. 39,120 The reason for this was<br />

that the techniques previously employed, i.e. resonance<br />

Raman and Stark effect spectroscopy, solely report on coherent<br />

states, like that of the localized 1 MLCT excited state, and not<br />

on delocalized states, like the initial, Franck–Condon excited<br />

state. Speculation of longer-lived <strong>charge</strong> hopping as a means<br />

of randomizing the ligand radical excited state is also not<br />

possible based on these observations, although the anisotropic<br />

results are still not fully understood. 35,121,122<br />

Subsequently, by femtosecond fluorescence upconversion it<br />

was shown that the lifetime of the 1 MLCT excited state of<br />

Ru(bpy) 3 2+ was 45 15 fs. As this measurement directly<br />

probes the spin of the electrons, this lifetime is that of the true<br />

singlet-to-triplet intersystem crossing to the vibrationally ‘hot’<br />

triplet manifold of states, Fig. 7–3. 123 This value agrees well<br />

<strong>with</strong> those obtained in water using time-resolved, femtosecond<br />

stimulated Raman spectroscopy and polychromatic, femtosecond<br />

fluorescence upconversion. 124,125 Spectral features<br />

lasting B300 fs and observed by femtosecond, magic-angle<br />

transient absorption spectroscopy were also assigned to relaxation<br />

of the <strong>charge</strong>-localized, 1 MLCT excited state of<br />

Ru(bpy) 3 2+ to the triplet-character thexi state. 126 As this<br />

method probes the absorption of states and not spin directly,<br />

the reported half-time (t 1/2 = B100 fs) provided an upper<br />

limit to the true intersystem-crossing lifetime. Additionally, it<br />

could be reporting on both intersystem crossing and vibrational<br />

cooling <strong>with</strong>in the manifold of triplet-character states,<br />

Fig. 7–3 and 7–4, respectively. Further evidence for such a<br />

Fig. 7 Lennard-Jones potential energy wells illustrating the relative<br />

electronic and vibrational energies and lifetimes for Ru(bpy)3 2+ . Both<br />

internal-conversion thermal relaxation (2) and intersystem crossing (3)<br />

occur in the sub-picosecond time scale while the lifetime of the thexi<br />

state (5) is up to a microsecond. Taken from Fig. 9 of ref. 129.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 121


process was obtained by employing similar measurements,<br />

however in addition to the sub-picosecond component, a<br />

higher energy (360 nm), longer-lifetime (B5 ps) transient<br />

feature was also present. 35 As the ligand radical has a rather<br />

high extinction coefficient here, this component was assigned<br />

to vibrational relaxation to form the thexi state. This vibrational–relaxation<br />

lifetime <strong>with</strong>in the manifold of states was<br />

shown to vary from B0.6 to 5.0 ps and be rather solvent<br />

dependent. 35,123,127,128 Using picosecond Kerr-gated, timeresolved<br />

resonance Raman spectroscopy the lifetime of this<br />

relaxation was shown to be B20 ps for homoleptic and heteroleptic<br />

Ru(bpy)3 2+ -based molecules of varying <strong>charge</strong>s and isotopic<br />

compositions and in a variety of solvents. 129 For<br />

comparison, N3 0 s 1 MLCT excited-state t 1/2 wasreportedtobe<br />

B30 fs and thermal relaxation <strong>with</strong>in its triplet-character manifold<br />

was found to occur <strong>with</strong> a B80 fs half-time when bound to<br />

mesoporous, nanocrystalline TiO 2 thin films. 107<br />

D Dye sensitization<br />

Some early dye sensitization studies employed Ru(bpy)3 2+<br />

dissolved in the external electrolyte. 130,131 However, it was<br />

soon found that anchoring the sensitizer to the semiconductor<br />

surface was a more practical approach. 132 Anchoring <strong>transition</strong>-<strong>metal</strong><br />

compounds to the TiO2 surface requires functional<br />

groups that can form strong bonds <strong>with</strong> the <strong>metal</strong>-oxide<br />

surface. Functional groups based on carboxylic acids, phosphonates,<br />

alcohols, amides, siloxanes, acetyl acetonates, and<br />

cyanides have all been tested. 72 The aforementioned dcb<br />

ligand <strong>with</strong> two carboxylic acid groups remains the most<br />

successful in terms of absolute efficiency in DSSCs. In 1979,<br />

Goodenough and co-workers first proposed that dehydrative<br />

coupling of carboxylic acid groups <strong>with</strong> surface titanols would<br />

result in the formation of ester-type linkages. 132 He suggested<br />

that the p* orbitals of the dcb ligand would promote rapid<br />

excited-state electron injection into the conduction band of<br />

TiO2 but not that of SnO2 or ZnO. The difference being one<br />

of symmetry as the TiO2 conduction band is comprised mainly<br />

of unfilled d orbitals where that of SnO2 and ZnO possess<br />

predominantly s-orbital character, Fig. 8(a)/(c). This latter<br />

suggestion now has some experimental verification. 104 Interestingly,<br />

the proposed coupling is optimal when the ester and<br />

bpy p-systems are co-planar, yet such a geometry is not found<br />

in the ground state due to unfavorable steric interactions. In<br />

crystal structures of the corresponding ethyl ester compound,<br />

the plane defined by the C–CQO of the ester group is skewed<br />

by 10–151 from the plane of the pyridine ring, Fig. 8(b). 133<br />

Furthermore, there is no measurable resonance enhancement<br />

of the symmetric COO stretching mode in Raman experiments<br />

further indicating that these groups are unconjugated in solution<br />

and when bound to TiO2. 134,135 However, upon MLCT excitation,<br />

the bpy ring is formally reduced by one electron and the<br />

ester group may twist. Persson et al. have shown computationally<br />

that the planar geometry enhances excited-state injection. 136<br />

Only under very acidic, non-aqueous conditions has<br />

evidence for an ester-type linkage been observed. 137 Physisorption<br />

through a solvation layer has been proposed by<br />

Hester and colleagues. 138 Under most conditions relevant to<br />

DSSCs, the predominant binding mode elucidated through IR<br />

Fig. 8 Orbital diagrams for ester-type binding to the surface of <strong>metal</strong><br />

oxides. (A) For TiO2, the overlap of the extended p system and the Ti<br />

3d orbitals are thought to aid in electron injection. (B) When carboxylates<br />

are rotated in such a way as to minimize orbital overlap, the<br />

injection yields are thought to suffer. (C) Similar effects are proposed<br />

for SnO2 as the Sn s orbitals have less efficient orbital mixing <strong>with</strong> the<br />

carboxylate p system. Adapted from Fig. 4 of ref. 132.<br />

studies is a carboxylate-type linkage; 137 unfortunately,<br />

the data does not allow for direct identification of the<br />

surface site(s) involved in the sensitizer–semiconductor<br />

bond. 132,134,137,139–141 Deacon and Phillips have tabulated<br />

vibrational data for <strong>metal</strong>-carboxylate compounds whose<br />

structures were determined crystallographically. 142 An empirical<br />

relation between the energy separation of the COO asymmetric<br />

and symmetric stretches and the carboxylate–<strong>metal</strong><br />

coordination mode was found. This same approach has been<br />

used to predict the carboxylate binding mode on the anatase<br />

TiO2 surface, presumably to Ti IV sites. 134,140,141 In agreement<br />

<strong>with</strong> theoretical studies, the analysis is most consistent <strong>with</strong> the<br />

carboxylate oxygens binding to separate Ti IV -<strong>metal</strong> centers.<br />

134,137,140,143 Such carboxylate linkages were observed<br />

even when the binding group was originally an ester, e.g. <strong>with</strong><br />

the deeb ligand which is 4,4 0 -(C2H5CO2)2-bpy. Therefore, we<br />

make no distinction between deeb and dcb throughout this<br />

review. Similarly, as the extent of deprotonation of sensitizers<br />

on the TiO2 surface is often unknown, the overall formal<br />

<strong>charge</strong> of semiconductor-bound sensitizers is often omitted.<br />

While <strong>transition</strong>-<strong>metal</strong> compounds based on dcb ligands<br />

remain the most successful for DSSCs, an important limitation<br />

is their poor stability in water. 144 Moderate stability has been<br />

reported in acidic electrolytes, but the sensitizers rapidly<br />

122 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


desorb when the pH is raised above pH B3.5. 145 In aqueous<br />

solutions, the most stable linkages appear to be those based on<br />

phosphonate groups. 144<br />

There now exists a large body of literature on the sensitization<br />

of TiO 2 by Fe II -, Ru II -, Os II - and Re I -polypyridyl compounds.<br />

146 There have also been some reports of sensitization<br />

by d 8 compounds based on Pt II , that also have MLCT-like<br />

excited states, and d 10 Cu I compounds. 147–149 Some of these<br />

results are highlighted in this review as alteration of the <strong>metal</strong><br />

center has, in some cases, provided insights into mechanistic<br />

details of dye sensitization.<br />

It is often tacitly assumed that the manifold of MLCT<br />

excited states observed in dilute solution or frozen glasses is<br />

unperturbed by the semiconductor surface. This assumption is<br />

often necessary as ultrafast injection precludes characterization<br />

of the excited state. However, as described in more detail<br />

below, the acceptor states in TiO 2 can be widely tuned in<br />

energy by controlling the concentration of potential-determining<br />

ions at the interface. With this approach and by utilizing<br />

sensitizers that are weak photoreductants, data on MLCT<br />

excited states anchored to TiO2 are now becoming available.<br />

One interesting finding is that the proximity of the sensitizers<br />

to one another on the surface affords efficient lateral energy<br />

<strong>transfer</strong> across the semiconductor surface. 150,151 Monte-Carlo<br />

simulations indicate a (30 ns) 1 energy <strong>transfer</strong> hopping rate<br />

constant at saturation surface coverage. 152 There is also evidence<br />

that the ligand-field states are destabilized upon surface binding.<br />

For example, compounds of the type [cis-Ru(bpy) 2(ina) 2] 2+ ,<br />

where ina is isonicotinic acid, are non-emissive in fluid solution<br />

<strong>with</strong> high quantum yields for photo-induced ligand loss, behavior<br />

that is expected for compounds <strong>with</strong> low-lying, ligand-field<br />

excited states. However, upon binding to MO2 (M = Ti or Zr)<br />

thin films, the compounds were found to be photoluminescent<br />

<strong>with</strong> temperature-dependent, excited-state lifetimes that were<br />

B50 ns at room temperature. 54 Both static and dynamic<br />

excited-state quenching were observed as the temperature was<br />

raised providing direct evidence that the intersystem-crossing<br />

quantum yield was temperature dependent and less than unity.<br />

Interestingly, when bound to TiO 2 thin films there was an<br />

inverse relation between the temperature and the quantum yield<br />

for photo-induced, interfacial electron injection, herein referred<br />

to as the ‘injection yield.’<br />

3. Photo-induced, interfacial <strong>charge</strong> separation<br />

The excited-state, interfacial-<strong>charge</strong>-separation mechanism<br />

shown in Fig. 1 is in fact only one of three mechanisms<br />

identified for electron injection. Said mechanisms differ<br />

by the state of the sensitizer and location of the electron<br />

that is <strong>transfer</strong>red to the semiconductor: (1) the excited<br />

state, i.e. [Ru III (bpy)2(dcb )] 2+ *; (2) the reduced state, i.e.<br />

[Ru II (bpy)2(dcb )] + ;or(3)via a molecule-to-particle <strong>charge</strong><br />

<strong>transfer</strong> event, i.e. Ru II –CN–Ti IV . An important variable<br />

for all of these sensitization mechanisms is the overlap of<br />

the molecular donor levels <strong>with</strong> the acceptor states of the<br />

semiconductor.<br />

Gerischer formulated a theory for excited-state injection<br />

into wide-bandgap semiconductors. 153–155 The rate of interfacial<br />

electron <strong>transfer</strong> at an electrode surface is proportional<br />

to the overlap of occupied donor excited states <strong>with</strong> unoccupied<br />

acceptor states:<br />

k inj B R k(E)D(E)W don(E) dE (6)<br />

where E is the energy, k(E) is the <strong>transfer</strong> frequency, D(E) is<br />

the density of unoccupied acceptor states (DOS) in the semiconductor,<br />

and W don(E) is the sensitizer donor distribution<br />

function. Fluctuations in the solvation of the sensitizer give<br />

rise to a distribution of excited-state energies. Gerischer<br />

defined the Gaussian donor and acceptor excited-state distri-<br />

bution functions, W(E):<br />

1<br />

WðEÞ ¼pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

exp<br />

4plkBT<br />

ðE EÞ 2<br />

!<br />

4lkBT<br />

where l is the reorganization energy of interfacial electron<br />

<strong>transfer</strong>, kB is Boltzmann’s constant, T is the temperature, and<br />

1E is the energy of the most probable solvation state. Thus, the<br />

rate constant of, and often the efficiency for, injection from the<br />

sensitizer are expected to depend critically on the overlap of<br />

the sensitizer excited-state distribution function <strong>with</strong> the<br />

semiconductor DOS.<br />

A Density of states in nanocrystalline TiO2 thin films<br />

used in DSSCs<br />

What are the density of unoccupied acceptor states, i.e. DOS,<br />

in nanocrystalline, anatase TiO2 thin films? This question<br />

remains somewhat unresolved. The classical method for determining<br />

these in the solid state is via photoelectron spectroscopy.<br />

Hagfeldt and co-workers have reported such data for a<br />

nanocrystalline TiO2 thin film sensitized <strong>with</strong> N3 in the<br />

presence and absence of Li + salts, Fig. 9(a). 156 This data<br />

shows a broad distribution of trap states centered at B1 eV<br />

below the energy of the conduction band edge (Ecb). However,<br />

it is well known that the flatband potentials of the semiconductors<br />

are very sensitive to environment. Therefore, the<br />

absolute and relative energies in vacuum may not be as<br />

relevant to a DSSC. In the field of photoelectrochemistry,<br />

the standard approach for determining the flatband potentials<br />

of semiconductor electrodes is Mott–Schottky analysis of<br />

capacitance data. 157 The analysis is based on the potentialdependent<br />

capacitance of a depletion layer at the semiconductor<br />

surface, behavior that is not likely observed for B20 nm<br />

anatase crystals that are expected to be fully depleted near<br />

kT. 18,158–165 Rothenberger and co-workers have proposed an<br />

accumulation-layer model to describe the potential distribution<br />

<strong>with</strong>in the TiO 2 particles at negative applied potentials. 166<br />

This model assumes that the band-edge positions remain fixed<br />

as the Fermi-energy is raised into accumulation conditions,<br />

behavior that has little literature precedence in electrolyte<br />

solutions. Nevertheless, the model provides the only literature<br />

estimates of Ecb available for these materials in organic and<br />

aqueous solvents <strong>with</strong> common electrolytes. 166–170 The literature<br />

values give the impression that the nanocrystalline TiO2<br />

thin films have a well-defined Ecb. Even if this is the case, there<br />

is a tremendous compilation of data supporting the notion<br />

that the acceptor states relevant to interfacial <strong>charge</strong> separation<br />

and recombination are more localized and are reduced<br />

more easily than literature E cb values indicate. 171–174<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 123<br />

ð7Þ


Fig. 9 (A) The density of acceptor states, DOS, for TiO2 thin film electrodes as measured by photoelectron spectroscopy and electrochemical methods<br />

(smallerplot).Thefigureswerescaledsoastoaligntheenergiesofthe(surface) deep trap states, exponential DOS near the conduction band, and<br />

conduction band edge; however, the energy differences among these states are dissimilar. Adapted from Fig. 1(b) of ref. 156 and Fig. 3 of ref. 171.<br />

(B) A diagram depicting the proposed energetic and spatial location of these same states as a function of their depth in a nanoparticle relative to the<br />

energy of the conduction band edge, Ecb, and the energy of the solution redox electrolyte, EF,redox. Adapted from Fig. 2(a) of ref. 171.<br />

Many electrochemical, photochemical, and spectroscopic<br />

studies have supported the suggestion that mesoporous, nanocrystalline<br />

TiO2 thin films possess a tailing of the DOS rather<br />

than an abrupt onset from an ideal Ecb. Determination of the<br />

precise form of these tailing states is non-trivial, however a<br />

novel computational method for determination of the absolute<br />

DOS distribution at zero Kelvin was recently reported by<br />

Bisquert and Zaban, and colleagues. 172,174 Although fundamentally<br />

important, the room temperature apparent DOS<br />

distribution is more relevant to the functioning DSSC. At room<br />

temperature, this distribution is thought to have an exponential<br />

dependence on the applied voltage as determined from electrochemical<br />

techniques where Fermi-level pinning was deduced<br />

to be negligible, Fig. 9(a), inset, 171,173–175 and recently by<br />

a spectroelectrochemical procedure. 176 Additionally, nonexponential<br />

kinetics for excited-state electron injection can be<br />

rationalized by invoking an exponential DOS at the TiO 2<br />

surface. 177–180 And by assuming said distribution is composed<br />

of bulk, intra-bandgap states, Fig. 9(b), diffusion of TiO 2<br />

electrons, TiO2(e )s,8 and dispersive recombination kinetics<br />

can be modeled satisfactorily by employing a multiple-trapping,<br />

continuous-time random walk model. 178,181–186 In addition, via<br />

these same techniques, Kavan et al., and many others since,<br />

have reported that TiO2 thin-film electrodes contain a relatively<br />

large population of deep, surface trap states at an energy<br />

located <strong>with</strong>in the bandgap and prior to a significant portion<br />

of the exponential distribution. 171,187–193 These states are believed<br />

to be unsaturated Ti IV surface states where oxygen<br />

vacancies reside. The energetics of such states were shown to<br />

be affected by surface chelation from various molecules due to<br />

the Lewis acidic and basic characteristics of the unsaturated<br />

Ti IV and surface-bound molecules, respectively. 188,189,192,194<br />

8 In much of the literature on nanocrystalline, anatase TiO 2, there is<br />

contention as to whether electrons in TiO2 are located in the diffuse<br />

conduction band, bulk exponential trap states, or deep surface trap<br />

states. For simplicity and clarity, collectively they will be denoted as<br />

TiO 2(e )s throughout this review. Although it is often implied that<br />

TiO 2(e )s are those located <strong>with</strong>in the exponential DOS, no distinction<br />

will be made unless it aids in understanding the desired point.<br />

The TiO2(e )s inferred from electrochemical measurements<br />

have spectroscopic signatures as well. As the indirect bandgap<br />

of anatase TiO2 is 3.2 eV, its ground-state UV-Vis absorption<br />

spectrum consists of a fundamental absorption edge at<br />

B385 nm and, in some cases, an Urbach tail at longer<br />

wavelengths. 195–198 The features observed for TiO 2(e )s in<br />

mesoporous, nanocrystalline TiO 2 (anatase) thin-film electrodes<br />

consist of a minor Burnstein–Moss shift, i.e. a blue shift in the<br />

fundamental absorption edge, 199,200 and a gradual rise in<br />

absorbance that tails to the near-IR, 201,202 and peaks at<br />

B1350 nm. 203 The extinction coefficient for the broad,<br />

featureless, visible-near-IR absorbance ranged from 640 to<br />

1300 M 1 cm 1 at 700 to 800 nm based on choice of solvent<br />

and electrolyte. 166,187,204,205 Similar features were present upon<br />

electrochemical bias of a single-crystal TiO2 (rutile) electrode<br />

to form TiO 2(e )s: a broad near-IR spectroscopic feature that<br />

peaked at 1500 nm. 206 As evidenced by spectroelectrochemical<br />

measurements, in addition to the current required to generate<br />

the ‘‘typical’’ TiO 2(e ) absorption features, an additional<br />

current pre-peak has been observed that is often largest in<br />

aqueous electrolyte, Fig. 10(a). 187,189,190 This has been<br />

ascribed to filling deep, surface trap states. It was determined<br />

that B12 of these surface states existed per 12 nm nanocrystallite<br />

and that they exhibited an absorption peak centered at<br />

B400 nm (e400 nm = B1900 M 1 cm 1 ), Fig. 10(b). 187<br />

Additionally, a new, broad absorption peak centered at<br />

B750 nm was observed (e 700nm = B2200–2800 M 1 cm 1 ),<br />

after passing 440 mC cm 2 (B100 TiO 2(e )/particle) in the<br />

presence of cations <strong>with</strong> large <strong>charge</strong>-to-radius ratios, i.e. Li + ,<br />

Na + , in acetonitrile or strongly basic aqueous electrolytes,<br />

Fig. 10(c). 205,207,208 This feature is indicative of small cation<br />

intercalation into the anatase lattice to form new<br />

phases. 156,190,209–215 Li + intercalation into highly reduced<br />

anatase TiO2 is known to form Li0.5TiO2 phases, 216,217 however<br />

such phases are not expected to be relevant to<br />

operational DSSCs.<br />

The TiO2(e ) states above are often described as shallow<br />

trap states and not entirely free conduction band electrons as<br />

their absorption would tail much farther into the IR, 218,219<br />

124 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 10 (A) A cyclic voltammogram of a TiO2 thin-film electrode in aqueous electrolyte. The large, reversible peak was indicative of filling and<br />

emptying the TiO2 DOS whereas the smaller pre-peak, present during the cathodic scan only, was assigned to the filling of deep trap states. Taken<br />

from Fig. 3(a) of ref. 187. (B) The absorption spectra of these biased electrodes illustrated the spectroscopic features associated <strong>with</strong> occupation of<br />

deep trap states, at 0.30 V and in bold, and formation of TiO2(e )s, at 0.70 V. Taken from Fig. 2 of ref. 187. (C) The absorption spectra of a<br />

thin film electrode in LiClO4 acetonitrile electrolyte biased to 1.50 V where formation of a new species, i.e. Li0.5TiO2 phases, was clearly evident<br />

near 750 nm. Taken from Fig. 3(a) of ref 205.<br />

they exhibit a sharp electron paramagnetic resonance (EPR)<br />

spectrum at 77 K, 220–222 and their apparent DOS follows an<br />

exponential distribution 156,171–175,177–180 <strong>with</strong> a non-ideality<br />

factor often greater than one. 171,178,183–185,223,224 An apparent<br />

exponential DOS distribution is also expected from theory for<br />

an ideal intrinsic semiconductor even though the actual underlying<br />

DOS distribution follows a power-law relationship <strong>with</strong><br />

energy. 225 However, the presence of a non-ideality factor<br />

unequal to unity is often attributed to a large concentration<br />

of trap states 218,219 Not<strong>with</strong>standing, using time-resolved<br />

infrared (TRIR) spectroscopy it was shown that the transient<br />

absorption features of TiO 2(e )s in TiO 2 and TiO 2–Pt colloids<br />

can be modeled as a function of the wavenumber to the<br />

1.5 power, indicative of free conduction band electrons.<br />

203,226 As electrons are thought to trap in TiO2 at<br />

coordinatively unsaturated Ti IV atoms <strong>with</strong>in a picosecond it<br />

was proposed that trapped electron thermalization to the<br />

conduction band may be possible at room temperature.<br />

A final comment <strong>with</strong> regard to the semiconductor DOS is<br />

that they are not singular material parameters. The most wellknown<br />

example is the nearly Nernstian shift, i.e. 59 mV/pH<br />

unit, in aqueous solution over the pH range H 0 = 8 to<br />

H = +23 due to protonation/deprotonation of surface titanol<br />

groups on TiO2. 166,169,227,228 It has also been known for quite<br />

some time that the flatband (and conduction band edge)<br />

potential of mesoporous, nanocrystalline TiO2 (anatase) can<br />

be widely tuned by the presence of cations in non-aqueous<br />

supporting electrolyte. This affect is greatest <strong>with</strong> cations<br />

possessing a large <strong>charge</strong>-to-radius ratio in the order Mg 2+<br />

4 Li + 4 Na + 4 K + 4 TBA + . 167,168 For example, Ecb has<br />

been reported to be 1.0 V vs. SCE ( 0.76 V vs. NHE 229 )in<br />

0.1 M LiClO 4 acetonitrile electrolyte and B 2.0 V ( 1.76 V)<br />

when Li + was replaced by TBA + . The direction of the bandedge<br />

shifts has been confirmed by excited-state quenching data<br />

described below. Interestingly, this same order has been<br />

observed for the equilibrium constants for cation adsorption<br />

onto TiO2 in aqueous solutions 230–233 and an electrolyte’s<br />

‘‘drying effect,’’ ionic association constant in aprotic solvents,<br />

and hydroxide association constant. 234 Although this shift<br />

is non-Nernstian, the behavior has been shown to be logarithmic<br />

in LiClO4 activity in acetonitrile and other aprotic<br />

mixed solvent systems. 167,168 Similar behavior was not observed<br />

in protic solvents hypothesized to be due to selective solvation<br />

of Li + by the protic solvent molecules. 167,168 In TBA + salts the<br />

flatband potential has been shown to depend logarithmically on<br />

the auto-ionization/autoprotolysis constant of the solvent. 167,168<br />

Thus, most likely, the large variations in Ecb (41V)inducedby<br />

the above ‘potential determining’ ions can be wholly explained<br />

by cation-coupled reduction potentials for TiO2 acceptor states,<br />

due to surface adsorption and/or intercalation into the anatase<br />

lattice. This same cation-dependent shift in E cb can be used to<br />

promote photo-induced electron injection from surface-bound<br />

sensitizers.<br />

B Ultrafast, excited-state electron injection<br />

After light absorption, the MLCT excited state of the sensitizer<br />

may inject an electron into the anatase nanocrystallite, a<br />

process also referred to as interfacial <strong>charge</strong> separation. For<br />

sensitizers like N3, light absorption formally promotes an<br />

electron from the <strong>metal</strong> center to a dcb ligand that is directly<br />

bound to the semiconductor surface. Therefore, excited-state<br />

<strong>charge</strong> separation occurs from the p* orbitals of the organic<br />

ligand to the acceptor states in TiO2, Fig. 8(b). There is now an<br />

overwhelming body of data that indicates that such <strong>charge</strong><br />

separation occurs on a femto- to pico-second time scale.<br />

Experimentally, ultrafast spectroscopists have all found that<br />

excited-state electron injection into TiO2 is non-exponential,<br />

behavior attributed to the surface heterogeneity of TiO2 and its<br />

DOS, distributions of sensitizer binding modes, strengths, and<br />

interactions, and multiple ultrafast injection processes occurring<br />

from various states in the thermal relaxation pathway, i.e.<br />

Franck–Condon singlet injection, internally converted singlet<br />

injection, intersystem crossing to the triplet state(s) followed by<br />

injection. This has been thoroughly reviewed for both organic<br />

and <strong>transition</strong>-<strong>metal</strong> coordination compounds bound to semiconductor<br />

<strong>metal</strong> oxides. 102,115,119 While the explanations given<br />

to rationalize the complex kinetics observed for excited-state<br />

injection for Ru II sensitizers are often reasonable, satisfactory<br />

mechanistic models are still lacking.<br />

It has been suggested that ultrafast, interfacial <strong>charge</strong> separation,<br />

following light absorption, does not always occur from the<br />

thexi state but rather often from the initial, Franck–Condon<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 125


excited state. Evidence for room-temperature injection occurring<br />

<strong>with</strong> a lifetime faster than a molecular vibration, i.e. kBT/h =<br />

B1.6 10 13 s = 160 fs, 235,236 eludes to this phenomenon. 102–119<br />

This would imply that injection is occurring before thermal<br />

relaxation of the molecular excited state.<br />

i Coherent, singlet injection. Willig and co-workers found<br />

that excited-state electron injection from N3* into TiO 2<br />

occurred in o25 fs under ultrahigh-vacuum conditions. 109<br />

The process therefore did not involve redistribution of vibrational<br />

excitation energy by exchange <strong>with</strong> phonons in the solid<br />

and thus was entirely different from the weak-electroniccoupling<br />

case of Marcus–Levich–Jortner–Gerischer-type electron<br />

<strong>transfer</strong>. 154,155,237–240 The finite reaction time for injection<br />

ruled out direct excitation of an electron from the Ru II -<strong>metal</strong><br />

center to the semiconductor, yet the sub-100 fs rise-time implied<br />

vibrational wave packet motion-induced electron <strong>transfer</strong>. A<br />

detailed analysis of theoretical and empirical results supporting<br />

these conclusions using a perylene sensitizer can be found<br />

elsewhere. 241–244 Briefly, clear, resolvable periodic beats,<br />

Fig. 11(a), were observable in the ultrafast transient signals<br />

consistent <strong>with</strong> coherent, singlet injection from the perylene<br />

singlet excited state. Fourier transform of these beats adequately<br />

reproduced the normal Raman modes of perylene,<br />

Fig. 11(b),(c). Thus, the oscillatory behavior was ascribed to<br />

pulsed electron <strong>transfer</strong> due to periodic surface crossing to the<br />

non-linear DOS in the semiconductor realized by vibrational<br />

wave-packet motion in the excited perylene, Fig. 11(d).<br />

The quantitative, ultrafast excited-state electron injection<br />

reported for N3/TiO 2 under ultrahigh-vacuum conditions was<br />

not always observed when the sensitized thin films were placed<br />

in organic solvents or electrolytes. Under such conditions,<br />

injection was non-exponential and occurred on the femtosecond<br />

to hundreds-of-picoseconds time scale. For N3 and<br />

porphyrin-based sensitizers, Durrant and co-workers found<br />

that a sum of three exponentials was required to fit the<br />

injection data, including an ultrafast o100 fs component.<br />

117,118 Interestingly, the rate constants for bpy- and<br />

porphyrin-based dyes were similar. These same authors later<br />

found that N719—the dianion salt of N3 <strong>with</strong> TBA + counterions—had<br />

a 30-fold slower rate of injection as compared to<br />

N3. 245 After performing multiple washings of the N3/TiO2<br />

films in neat ethanol the injection rates were similar to that of<br />

N719/TiO2 films. It was suggested that the labile protons from<br />

the carboxylic acid binding groups of N3 had lowered the Ecb<br />

and promoted more favorable energetics for injection. To<br />

control this variable Lian and co-workers pre-treated<br />

N3/TiO 2 thin films for one day in aqueous buffer solutions<br />

at pH 2, 4, 6, or 8. 103 After removing weakly bound and<br />

desorbed sensitizers, the biphasic kinetics and injection yields<br />

were found to be pH dependent. As the pH was raised from<br />

2 to 8, there was a decrease in the rate of the slower component<br />

to injection, the ratio of the faster-to-slower components to<br />

injection, and the injection yield. Such behavior is consistent<br />

<strong>with</strong> the expected Nernstian shift of Ecb towards the vacuum<br />

level as the pH is raised.<br />

Gra¨tzel and co-workers reported that the slower picosecond<br />

components for excited-state electron injection could be<br />

by employing a low concentration or sonicated dying solution<br />

or a lower surface-coverage thin film. 246,247 Under such conditions,<br />

only an ultrafast component (o20 fs) for injection<br />

remained. In support of this, Piotrowiak and co-workers found<br />

that dialysis of sensitized TiO2 colloids resulted in much shorter<br />

excited-state lifetimes as measured by time-correlated single<br />

photon counting. 248 However, in this case multi-exponential<br />

kinetics were required to adequately fit the observed data.<br />

Lian and co-workers found that excited-state electron injection<br />

into TiO2 was biphasic for three [cis-Ru(dcb)2(X)2] 0,0,2+<br />

Fig. 11 (A) An ultrafast, time-resolved, single-wavelength absorption difference spectrum for perylene/TiO2 displaying periodic beats. The<br />

Fourier transform, (B), of the periodic beats, (inset), effectively reproduced the normal modes of perylene, (C). (D) A schematic depicting<br />

the model used to rationalize the empirical data; periodic crossing of the molecular vibrational wavepacket <strong>with</strong> the TiO2 DOS. Taken from<br />

Fig. 5 and 10, respectively of ref. 244.<br />

126 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


compounds (X = SCN ,X=NC , or (X)2 = dcb) and fit a<br />

two-state model. 103 The rate of the slower component was<br />

directly related to the sensitizer excited-state reduction potential<br />

while the relative magnitude showed the opposite trend.<br />

No noticeable changes were apparent for the fast component<br />

<strong>with</strong>in the time resolution of the measurement, i.e. B200 fs.<br />

With Re(dcb)CO3Cl/TiO2 it was suggested that ultrafast<br />

injection (o50 fs) was from a vibrationally ‘hot’ state. 103–106<br />

As measured by femtosecond TRIR spectroscopy, the CQO<br />

stretching band in the excited state red-shifted by 10 cm 1 over<br />

10 ps. The difference in rate constants implied that injection<br />

occurred before thermal electron relaxation and reorganization<br />

of the inner-sphere ligand environment. This same group<br />

reported the injection dependence for carboxylic acid versus<br />

phosphonic acid linkers <strong>with</strong> Re(dmb-X 2)CO 3Cl sensitizers,<br />

where dmb is 4,4 0 -dimethyl-bpy (X = COOH or PO 3H 2). 116<br />

The sensitizer <strong>with</strong> X = PO 3H 2 resulted in faster injection which<br />

was in conflict <strong>with</strong> previous findings employing organic sensitizers.<br />

108 However, the experimental data was supported by DFT<br />

calculations on the anionic versions of the sensitizers showing<br />

that there was a stronger electronic coupling between the dmb-X2<br />

and Ti IV -<strong>metal</strong> centers when bound through phosphonate<br />

linkages. 116 Additionally, solvent-dependent injection rates<br />

were studied using Re(dcb)CO3Cl sensitizers. 249 It was found<br />

that the rate of the slow, picosecond component for injection<br />

decreased in the order water (pH 2) 4 MeOH E EtOH 4 water<br />

(pH 8) 4 DMF which could be expected based on the proposed<br />

E cb for TiO 2 and electron <strong>transfer</strong> in the Marcus normal region.<br />

However, changes were not as large as expected due to trace<br />

water adsorbate whose presence was verified by FTIR.<br />

McCusker and co-workers found excitation wavelengthdependent,<br />

tri-exponential kinetics for N3/TiO2, cis-<br />

Ru(dcb)2(CN)2/TiO2, and their osmium analogues. 112 For<br />

the Ru II -based sensitizers, excitation at shorter wavelengths<br />

resulted in a larger amplitude femtosecond component, assigned<br />

to electron injection from the 1 MLCT excited state, and<br />

thus a smaller picosecond amplitude, assigned to injection<br />

from the 3 MLCT excited state. On the picosecond-time scale,<br />

3 MLCT components were much more dominant for osmium<br />

analogues where the spin–orbit coupling was larger. For all<br />

sensitizers studied, the rate of the picosecond component was<br />

found to be directly related to the sensitizer excited-state<br />

reduction potential, E o (Ru III/II* ), consistent <strong>with</strong> electron injection<br />

from the thexi state.<br />

At about the same time, Sundstro¨m and co-workers reported<br />

stimulated emission from the initially formed, singlet<br />

excited state of N3 <strong>with</strong> a B70 and B30 fs half rise-time<br />

in solution and on TiO2, respectively. 107,111 These time<br />

scales were similar to those measured by femtosecond transient<br />

absorption spectroscopy for intersystem crossing,<br />

Fig. 12(b). 107,111,114 The branching ratio for electron injection<br />

from the 1 MLCT state and intersystem crossing to the 3 MLCT<br />

state resulted in time constants of B50 and B75 fs for each<br />

process, respectively. Excitation into the low-energy shoulder<br />

of N3’s absorption spectrum was shown to directly populate<br />

N3’s 3 MLCT manifold both in solution (t = B70 fs) and on<br />

TiO 2. It was also shown that injection became slower and less<br />

efficient, i.e. from B100% to B50%, as the excitation light<br />

was shifted towards longer wavelengths, Fig. 12(b). This was<br />

postulated to be due to injection from a manifold of 3 MLCT<br />

excited states. Similar findings have been observed for<br />

[Ru(bpy)2(dcb)] 2+ on SnO2 but resulting in slightly larger<br />

half-times, 250 behavior that is consistent <strong>with</strong> Goodenough’s<br />

hypothesis. 132 These same authors found that by varying the<br />

method of TiO2 film preparation, both rate constants for the<br />

biphasic injection kinetics for N3* into TiO 2 were directly<br />

related to the degree of TiO 2 crystallinity. 110 Similar effects<br />

have been observed <strong>with</strong> organic sensitizers. 251,252<br />

As mentioned previously, spin arguments <strong>with</strong> Ru II and<br />

Os II sensitizers are complicated by spin–orbit coupling that<br />

effectively mixes the spin states, no longer making spin a good<br />

quantum number. With organic sensitizers this is not the case<br />

and well-defined singlet and triplet states have been shown to<br />

sensitize TiO2. With a Ti IV -phthalocyanine sensitizer anchored<br />

to TiO2 via an axial 3,4,-dihydroxybenzoic acid ligand, wavelength-dependent<br />

injection yields were apparent. 253 Although<br />

such behavior could have been attributed to ‘hot’ injection,<br />

this was not thought to be the case here. The rate constants<br />

for excited-state injection from the equilibrated singlet and<br />

triplet excited states were proposed to be different. This stateselective<br />

injection was assigned to efficient kinetic competition<br />

between injection from the S2 excited state (from excitation<br />

Fig. 12 (A) Picosecond transient absorption difference spectra for N3/TiO 2 where changes due to 3 MLCT excited-state injection are noted. Taken<br />

from Fig. 2 of ref. 107. (B) Time-resolved, single-wavelength absorption difference spectra for N3 and N3/TiO2 demonstrating relaxation <strong>with</strong>in<br />

the triplet-character manifold of states on the picosecond time scale. Excitation wavelengths are indicated on the figure. Taken from Fig. 4 of<br />

ref. 114. (C) A schematic depicting the possible interfacial, excited-state processes: (a) ultrafast, ‘hot’ injection; (b) intersystem crossing;<br />

(c) vibrational relaxation; and (d) slower thexi-state injection. Taken from Fig. 9 of ref. 111.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 127


into the Soret band) and internal conversion/vibrational relaxation<br />

of this state to the lower lying S1 state (that can be<br />

directly populated <strong>with</strong> excitation into the Q bands).<br />

ii TiO 2 DOS and quasi-Fermi-level dependence. Another<br />

school of thought is that the multiphasic, excited-state electroninjection<br />

kinetics do not result solely from heterogeneity of the<br />

sensitizer but also reflect differences in the TiO2 DOS, Fig. 13.<br />

There is growing evidence that a well-defined Ecb is not<br />

relevant to excited-state injection for these sensitized nanocrystalline<br />

thin films. The previously mentioned slower injection<br />

for N719 over N3, is thought to result from proton<br />

adsorption-induced shifts in the DOS. 245 Interestingly, even<br />

though injection was observed to be slower after excitation of<br />

N719/TiO2, the energy conversion efficiency was found to be<br />

higher. 177 This occurs because proton-induced shifting of the<br />

DOS is positive on an electrochemical scale and can thus lower<br />

the V oc. Keep in mind that due to the long-lived nature of the<br />

MLCT excited states, a quantitative injection yield could<br />

occur even if injection dynamics were slowed to the few<br />

nanosecond time scale. Indeed, it was shown that multiphasic<br />

injection half-times for N719 and N3 were 12 and 0.4 ps,<br />

respectively, <strong>with</strong> over an order-of-magnitude slower <strong>charge</strong>separation<br />

dynamics for N719. The injection kinetic data fit a<br />

model employing an exponentially increasing DOS, which is<br />

apparent elsewhere in the literature, 171,173–175,178–180 and activationless<br />

excited-stated electron <strong>transfer</strong> from a Gaussian<br />

distribution of energy offsets. Monte Carlo numerical simulations<br />

178,179 were in excellent agreement <strong>with</strong> the empirical<br />

reaction dynamics. It was also found that the excited-state<br />

injection kinetics for N719 were over 20 times slower in a full<br />

DSSC versus an inert electrolyte. However, the complete<br />

DSSC still showed excellent photovoltaic performance, due<br />

to the sluggish <strong>charge</strong> recombination kinetics and minimal<br />

Fig. 13 A Gerischer Diagram illustrating excited-state electron injection<br />

from surface-bound sensitizers into the DOS of the TiO 2<br />

nanocrystallites. E is the electrochemical potential of the conduction<br />

band edge (E cb), of the deep trap states (E T), and of the sensitizer at<br />

standard-state conditions (E 0 (A/D) and E 0 (A/D*), for the ground and<br />

thexi states, respectively). D(E) is the TiO 2 DOS, W don(E) and<br />

Wdon*(E) are the sensitizer donor distribution functions of the ground<br />

and thexi states, W acc(E) and W acc*(E) are the sensitizer acceptor<br />

distribution functions, and l is the reorganization energy. Adapted<br />

from Fig. 3 of ref. 171 and Fig. 3(a) of ref. 119.<br />

‘kinetic redundancy,’ where the time scale for injection was<br />

sufficiently less than the excited-state lifetime but not by an<br />

excessive amount.<br />

We have shown that the excited state of Ru(bpy) 2(dcb)/TiO2 thin films immersed in acetonitrile exhibit both static and<br />

dynamic quenching when Li + is introduced into solution. 254<br />

This was ascribed to photo-induced electron injection into<br />

TiO2 acceptor states where said states become thermodynamically<br />

accessible due to the positive cation-induced shift of<br />

the DOS. This was further supported by the monotonic and<br />

somewhat linear increase in both PLI/PLIo and injection yield<br />

<strong>with</strong> the logarithmic concentration of Li + (PLI is photoluminescence<br />

intensity). The trend for such behavior was linear in<br />

the <strong>charge</strong>-to-radius ratio of the 2 mM cation employed in the<br />

order Ca 2+ 4 Ba 2+ E Sr 2+ 4 Li + 4 Na + 4 K + 4 Rb +<br />

E Cs + E TBA + , and smallest for neat acetonitrile.<br />

As mentioned previously, in the absence of such external<br />

cations the flatband potential has been reported to scale<br />

logarithmically <strong>with</strong> the solution auto-ionization constant. 167,168<br />

This implies that proton activity determines the potential of the<br />

TiO2 DOS in these neat solvents. Thus, a strategy to introduce<br />

cations <strong>with</strong> a large <strong>charge</strong>-to-radius ratio into non-aqueous<br />

electrolytes was employed: acid- and base-pretreatment of TiO2<br />

thin films using H2SO4, HCl,orHClO4and NaOH, respectively.<br />

137 It was shown that when [Ru(bpy) 2(deeb)] 2+ sensitizers<br />

were bound to acid pre-treated TiO2 films they bound as the acid<br />

form, i.e. –COOH, and injected electrons much better than base<br />

pre-treated films, which bound as the carboxylate form, i.e.<br />

–COO . In fact, injection yields in neat acetonitrile and IPCEs<br />

in TBAI/I2 electrolyte were o10% for pH 43 pre-treatment<br />

while for pH o2.5 pre-treatment injection yields were 480%.<br />

(It is of note that the point-of-zero <strong>charge</strong> of TiO2 is B5–6 255–258<br />

while the pKa of the sensitizer carboxylic acid groups are 1.75 and<br />

2.80.) 259 Upon addition of LiClO4 to base pre-treated films,<br />

injection yields increased significantly; for acid pre-treated<br />

films, most of the dyes desorbed.<br />

All of the previously described photo-induced electron<br />

injection studies were performed on equilibrated systems<br />

under open-circuit conditions. It is important to quantify<br />

interfacial <strong>charge</strong> separation under short-circuit conditions—when<br />

the system is initially at a steady state—and to<br />

specifically quantify the effect(s) TiO2(e )s have on excitedstate<br />

injection. The first such report studied the biasdependence<br />

on the injection yield from a photoexcited<br />

Ru(dcb)3/TiO2 thin-film electrode in a pH 3, 0.2 M LiClO4<br />

aqueous electrolyte. 158 Upon reverse bias or near open-circuit<br />

conditions, the injection yield was essentially unity. However,<br />

as the electrode was biased in the forward direction, closer to<br />

the operational power point of the electrode, the injection<br />

yield dropped to B0.5, Fig. 14. This was ascribed to the filling<br />

of the DOS in TiO2 leading to decreased injection and an<br />

increase in PLI. Similar behavior was observed on sensitized<br />

SnO2 electrodes. 260 A complication in these studies is that<br />

forward bias can result in desorption of the sensitizers from<br />

the semiconductor surface, which will by itself lower injection<br />

yields and increase the PLI. 261<br />

A seven-fold increase in the half-time for excited-state<br />

electron injection from fully deprotonated N3/TiO2 in acetonitrile<br />

was obtained by omission of Li + from the solution. 262<br />

128 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 14 The quantum yield of excited-state electron injection from<br />

surface-bound Ru(dcb) 3 2+ into TiO2 as a function of electrochemical<br />

applied bias. Inset: The photoluminescence spectra of Ru(dcb)3/TiO2<br />

thin film electrodes at the indicated potentials. Taken from Fig. 8 of<br />

ref. 158.<br />

Biasing the sensitized electrode to 700 mV vs. Ag/AgCl, the<br />

most negative bias where desorption/degradation did not occur,<br />

resulted in the same injection yield on the longest time scales<br />

studied, 600 ps, but <strong>with</strong> significant attenuation of the<br />

fast component to injection. The half-times for injection were<br />

25-fold slower at this applied bias and could be modeled by<br />

non-adiabatic electron <strong>transfer</strong> theory where, prior to injection,<br />

thermal equilibrium of the excited state was assumed. Since up to<br />

40% of the injection occurred on the sub-molecular vibration<br />

time scale, i.e. B160 fs, the injection kinetics were most likely<br />

modeled under conditions where the assumption was valid.<br />

iii Distance dependence. The multiphasic character and<br />

picosecond dynamics of excited-state electron injection into<br />

TiO 2 alludes to the idea that at least some injection is occurring<br />

from a thexi state. This state, which can be described by a<br />

Boltzmann population, may exhibit behavior typical of thermal<br />

electron <strong>transfer</strong> and/or electron tunneling. The latter is clearly<br />

evident by temperature- and distance-dependent studies. At low<br />

temperatures a constant, nonzero rate for injection may persist<br />

while the room-temperature injection rate constants ought to<br />

exhibit an exponential dependence on distance:<br />

k = koexp[ bx] (8)<br />

where b is the dampening factor. A dampening factor,<br />

b = B1.0 A˚ 1 , is often indicative of saturated-hydrocarbon,<br />

through-bond superexchange tunneling behavior; 263–266 generally,<br />

larger values imply at least partial through-space<br />

character while smaller ones are associated <strong>with</strong> tunneling<br />

through conjugated p systems. 266<br />

An early study demonstrated that efficient excited-state<br />

electron injection did occur from sensitizers of the general<br />

type Ru(dmb)2(L) 2+ , where L contained unconjugated<br />

–(CH2)x– linkers between the Ru-chelating bpy moiety and<br />

one carboxylic acid group. 72 A more systematic study was<br />

later reported using three Re(bpy(CH2)2n(COOH)2)CO3Cl<br />

(n = 0, 1, 3) sensitizers where it was shown that ultrafast<br />

injection into TiO 2 did not occur when electronic coupling<br />

between the surface-bound ligand and the TiO 2 surface was<br />

removed by unconjugated methylene spacers, i.e. when n =1<br />

or 3. 104,105 For the same two sensitizers, the slower picosecond<br />

injection process could be successfully fit to a stretched<br />

exponential and the distance dependence of the injection rate<br />

could be qualitatively modeled by eqn (8) using b =1.2foreach<br />

C–C bond, indicative of nonadiabatic electron <strong>transfer</strong>. The<br />

4200-fold increase in injection rate from n =1ton =0could<br />

not be fit to such a model and was explained as adiabatic electron<br />

<strong>transfer</strong> due to a greatly increased strong electronic coupling<br />

from the lack of an unconjugated spacer moiety, Fig. 15.<br />

Detailed comparison of the n = 0 <strong>with</strong> the n = 1 or 3 compounds<br />

were complicated by the fact that the n = 0 compound had<br />

significantly different photophysical and redox properties.<br />

The distance dependence of excited-state electron injection<br />

was also explored using Ru II sensitizers that contained a bpy<br />

ligand derivatized <strong>with</strong> a conjugated oligo(xylylene) linker and<br />

bound to TiO2 via an ethynylcarboxyphenyl group. 267 A series<br />

of three compounds, <strong>with</strong> zero, one, or two linkers, was<br />

employed. A mere two-fold difference in injection rate constant<br />

was inferred by the difference in integrated PL spectra of<br />

the dyes in solution and on TiO2. The lack of the expected<br />

large differences was proposed to result from the flexibility of<br />

the one-carboxyl sensitizer attachment, that allowed proximity<br />

of the Ru II -<strong>metal</strong> center and the TiO2 surface in all three cases.<br />

In a related study, sub-picosecond injection was observed for<br />

2+<br />

tripodal, Ru(bpy) 3 -based sensitizers <strong>with</strong> an oligo(phenyleneethynylene)<br />

linker covalently bound to a tricarboxyphenyladamantane<br />

base calculated to have Ru–TiO2 distances over<br />

24 A˚ . 268–270 This study did not systematically show that rates<br />

vary <strong>with</strong> distance but did provide strong evidence that the<br />

distance dependence on injection rate was not large.<br />

With three phosphonated, ‘black dye’-like compounds of<br />

the form [Ru(40-PO3(Ph)n-tpy)(NCS)3] 3<br />

(n = 0, 1, 2) the<br />

distance-dependence of excited-state electron injection<br />

through conjugated linkers was studied. 271 Femtosecond<br />

pump–probe transient absorption measurements revealed that<br />

Fig. 15 Time-resolved, single-wavelength absorption difference spectra<br />

for Re(bpy(CH2)2n(COOH)2)CO3Cl/TiO2 (n = 0, 1, 3) illustrating<br />

that the rate of injection was inversely related to n. Taken from Fig. 9<br />

of ref. 104.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 129


Fig. 16 (A) A diagram of a TiO 2/Al 2O 3 core-shell nanoparticle. (B) Time-resolved, single-wavelength absorption difference spectra for<br />

Ru(4 0 -PO3 2 -tpy)(NCS)3/TiO2 thin films illustrating that the rate of injection was inversely related to the size of the Al2O3 overlayer. Al2O3<br />

overlayer thickness in nanometers are shown. Taken from Fig. 5 and 6, respectively, of ref. 271.<br />

the rate of each phase of an observed biphasic injection<br />

process was dependent on distance. The fast picosecond<br />

component fit nicely to an exponential distance-dependent<br />

model, eqn (8), <strong>with</strong> dampening factor, b = 0.19 A ˚ 1 , while<br />

the slower component for injection was assumed to be due to<br />

injection from loosely bound or aggregated dyes. As this<br />

dampening factor was much smaller than typical values obtained<br />

for donor–bridge–acceptor systems in solution, it was<br />

proposed that nuclear reorganization played a negligible role<br />

in injection, a hypothesis supported by DFT calculations.<br />

The distance-dependence was investigated by yet another<br />

means using the sensitizer lacking phenylene bridges, i.e. n =0;<br />

a core-shell architecture 272–276 was employed <strong>with</strong> Al 2O 3 shells<br />

varying from 0.6–6 nm in thickness conformally deposited on<br />

TiO2 prior to thin film preparation, Fig. 16(a). 271 The insulating<br />

shell required tunneling for almost all excited-state injection.<br />

As tunneling is not only a factor of distance but barrier<br />

height as well, this architecture allowed solely the distance to<br />

be altered. Neglecting ultrafast injection, which was assumed<br />

to be from dyes adsorbed directly onto TiO2 from small holes<br />

in the Al 2O 3, it was shown that the picosecond biphasic nature<br />

of injection resulted in b = 0.11 A˚ 1 and 0.04 A˚ 1 for the fast<br />

and slow components, respectively, Fig. 16(b). As the barrier<br />

to the conduction band of bulk, crystalline Al 2O 3 is very large,<br />

dampening factors over an order-of-magnitude larger were<br />

expected. It was proposed that the electronic structure of thin<br />

alumina layers differed from that of bulk Al2O3. 277<br />

Using three rigid-rod, Ru(bpy)3 2+ -based compounds containing<br />

a conjugated bpy ligand derivatized <strong>with</strong> an oligo-<br />

(phenyleneethynylene) linker and anchored to TiO2 via a<br />

dicarboxyphenyl group the distance dependence of excitedstate<br />

electron injection was studied, Fig. 17(a). 278 It was found<br />

that a monotonic decrease in injection rate occurred as the<br />

number of linkers was increased. However, this dependence<br />

only resulted in a dampening factor, b = 0.04 A˚ 1 , for both<br />

the slow and fast picosecond components, whereas a similar<br />

study on SnO2 resulted in a value of B0.8 A ˚ 1 , 279 and<br />

theoretical values were 40.4 A ˚ 1 . Although this small distance<br />

dependence agrees rather well <strong>with</strong> the conclusions from<br />

the phosphonated, ‘black dye’-like compounds, these results<br />

were further complicated by the lack of an expected similar<br />

trend in injection yields, where the middle-length spacer was<br />

found to inject best, Fig. 17(b).<br />

Fig. 17 (A) A diagram of a rigid-rod, Ru(bpy) 3 2+ -based sensitizer<br />

bound to a TiO2 nanocrystallite. (B) Time-resolved, single-wavelength<br />

absorption difference spectra of these TiO 2-bound sensitizers containing<br />

rods of oligo(phenyleneethynylene) linkers (n = 1, 2, 3). Although<br />

the injection yields were not distance-dependent, the rates were<br />

inversely related to n. Taken from cover artwork and Fig. 3A,<br />

respectively, of ref. 278.<br />

130 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


The observation of efficient and rapid excited-state electron<br />

injection through saturated and unsaturated spacers raises the<br />

question of whether the MLCT excited state need be localized<br />

on a ligand that is directly attached to the semiconductor<br />

surface. In other words, could the surface linker be on a nonchromophoric<br />

ligand? An early test of this was performed <strong>with</strong><br />

a bi<strong>metal</strong>lic Re I (dcb)CO3–L–Ru II (bpy)2(CN) (L = CN or<br />

NC ) compound. 280 Long-wavelength excitation selectively<br />

promoted an electron from the Ru II -<strong>metal</strong> center to a bpy<br />

ligand that was not anchored to the semiconductor surface, yet<br />

still resulted in a large photocurrent in regenerative DSSCs.<br />

The dcb ligand is structurally the same as two ina ligands<br />

connected in the 2 and 2 0 positions. The extra covalent bond in<br />

the dcb ligand increases the overall conjugation and thus<br />

lowers its LUMO energy. Using a comparative study of two<br />

heteroleptic Ru II compounds, one <strong>with</strong> a dcb ligand and the<br />

other <strong>with</strong> two ina ligands, the effect of remote versus adjacent<br />

excited-state electron injection was directly studied, Fig. 18. 281<br />

Both compounds exhibited a similar pH-dependent injection<br />

at pH 42 even though the thexi state of the latter compound<br />

contained an electron localized on a ligand that was not bound<br />

to the TiO2 surface. The observations of efficient injection<br />

from sensitizers <strong>with</strong> an ina ligand has been observed for<br />

Re(bpy)CO3(ina) + as well. 282<br />

The ina ligand, and substituted analogues, can also be<br />

coordinated to axial sites in porphyrinic macrocycles. A Ru II -<br />

phthalocyanine sensitizer <strong>with</strong> axial 3,4-dicarboxylic acidpyridine<br />

was employed. 283 Thepyridinederivativeallowedfor<br />

surface binding of the sensitizer to TiO2 however the major near<br />

IR–visible light absorption features were due to intraligand<br />

p - p* <strong>transition</strong>s that were localized on the phthalocyaninato<br />

ligand. Quasi-monochromatic light excitation resulted in excitedstate<br />

electron injection into TiO2 <strong>with</strong> a maximum IPCE 4 60%,<br />

due entirely to remote injection. Similar remote injection results<br />

have been obtained for similar p - p* <strong>transition</strong> molecules:<br />

a Ti IV phthalocyanine <strong>with</strong> a 3,4-dihydroxybenzoic acid<br />

surface-binding ligand and other Ru II phthalocyanines <strong>with</strong> a<br />

4-carboxylic acid-pyridine surface-binding ligand. 253,284,285<br />

iv ‘Hot’ injection. It is somewhat surprising that the photoinduced<br />

ultrafast electron injection observed by so many<br />

Fig. 18 A schematic illustrating two different injection schemes<br />

depending on the surface-bound ligands: (A) Remote excitedstate<br />

injection pathway for cis-Ru(dpp) 2(ina) 2/TiO 2, where dpp is<br />

4,7-diphenylphenanthroline, due to excited-state localization on a<br />

dpp ligand. (B) Adjacent excited-state injection pathway for<br />

Ru(dpp)2(dcb)/TiO2 as the excited state is localized on the surfacebound<br />

dcb ligand. Taken from cover artwork of ref. 281.<br />

authors is not manifest in operational DSSCs. One might<br />

anticipate that blue photons would give rise to higher photocurrents<br />

than would red ones due to the stronger reducing<br />

power of the Franck–Condon excited states in the former case.<br />

In other words, the absorptance and photocurrent action<br />

spectra would deviate significantly from each other <strong>with</strong> much<br />

less photocurrent at long wavelengths than would be expected<br />

based on the fraction of light absorbed. A possible reason why<br />

such behavior is not commonly observed is that the presence<br />

of the redox mediator (typically 0.5 M LiI/0.05 M I2), slows<br />

injection. Keep in mind that most of the interfacial<br />

<strong>charge</strong>-separation data described above was obtained in inert<br />

electrolytes, solvent, or vacuum. The little data available for<br />

excited-state injection in the presence of the I3 /I redox<br />

mediator indicates, in fact, that <strong>charge</strong> separation is still<br />

quantitative but that the kinetics are significantly altered. 177<br />

Another possibility is that there is quantitative injection from<br />

the sum of upper—i.e. Franck–Condon, vibrationally ‘hot,’<br />

etc.—and thermally relaxed excited states so that the photocurrent<br />

action spectrum simply does not report on the ultrafast<br />

processes.<br />

In addition to the ultrafast, ‘hot’ injection observed in high<br />

vacuum, solvents, and electrolytes, there are in fact a few<br />

literature observations that indicate that photo-induced ‘hot’electron<br />

injection occurs in DSSCs. The first example was <strong>with</strong><br />

cis-Fe(dcb) 2(CN) 2/TiO2 that exhibited a band-selective photocurrent<br />

action spectrum. 286 Although the photocurrent efficiency<br />

was poor at all excitation wavelengths (o10%), there<br />

was over a five-fold increase in the absorbed-photon-tocurrent<br />

efficiency (APCE), or internal quantum efficiency, when<br />

the higher-energy absorption band was excited. This behavior<br />

was attributed to band-selective injection yields occurring<br />

when injection was not from the thexi state. Thus, this further<br />

supported the interfacial, ‘hot’-injection mechanism. Similar<br />

behavior was reported for cis-Ru(dcbq)2(NCS)2/TiO2, where<br />

dcbq is 4,40-dicarboxylic acid-2,20-biquinoline. 287<br />

It was<br />

shown that the APCE was wavelength-dependent <strong>with</strong> the<br />

lower-energy band exhibiting smaller APCEs and rationalized<br />

as injection from this band being thermodynamically unfavorable.<br />

This is indicative of ‘hot’-injection processes from both<br />

bands <strong>with</strong> additional thexi-state injection from the higher<br />

energy band only. The hypothesis was tested <strong>with</strong> SnO2 thin<br />

films, whose Ecb is B500 mV more positive than that of TiO2.<br />

The photocurrent action spectrum better traced the absorptance<br />

spectrum as the APCE was no longer wavelength<br />

dependent. The third example we are aware of was reported<br />

by Heimer and co-workers who also observed APCEs that<br />

were wavelength dependent on TiO 2, but not on SnO 2, <strong>with</strong><br />

N3-like derivatives <strong>with</strong> the carboxylic acid groups present in<br />

the 5 and 5 0 positions. 288<br />

An alternative approach for quantifying ‘hot’-electron injection<br />

is to measure the injection yields as a function of the<br />

excitation wavelength. Moser and Gra¨tzel first reported<br />

studies of this type. 289 For N3/TiO2, the nanosecond injection<br />

yields were quantitative and wavelength independent.<br />

However, when N3 was anchored to Nb2O5, who’s Ecb<br />

is more difficult to reduce, wavelength-dependent injection<br />

yields were observed that decreased as the excitation<br />

wavelength increased. When the sensitizer was changed to<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 131


cis-Ru(2,6-bis(1 0 -methylbenzimidazol-2 0 -yl)pyridine)(dcbq)-<br />

(NCS)2, the injection yields were found to be strongly<br />

wavelength dependent. The dcbq ligands have low-lying p*<br />

LUMO energies, too low for the thexi state of the sensitizer<br />

to inject an electron into TiO 2. Thus, ‘hot’ injection from<br />

an upper vibrational excited state was, <strong>with</strong>out-a-doubt,<br />

occurring. Therefore, wavelength-dependent injection yields<br />

measured on nanosecond time scales are a signature of<br />

injection from upper, or ‘hot’, excited states.<br />

We have recently observed similar behavior <strong>with</strong> Ru II -ammine<br />

sensitizers. 290 These compounds have low-lying, ligandfield<br />

states and, in some regards, have excited-state properties<br />

more similar to Fe II compounds than Ru(bpy)3 2+ . Wavelength-dependent<br />

injection yields were observed for<br />

Ru(NH 3) 5(ina)/TiO 2 <strong>with</strong> blue-light excitation giving almost<br />

twice the injection yield as <strong>with</strong> green light, Fig. 19. Interestingly,<br />

<strong>with</strong> the tetra-ammine compounds, Ru(NH 3) 4(dcb)/<br />

TiO2, the injection yields were also wavelength dependent<br />

and were sensitive to isotopic substitution of the ammines.<br />

The injection yields were about 30% larger for ND3 than NH3<br />

at all excitation wavelengths studied, data that is consistent<br />

<strong>with</strong> ‘hot’-electron injection.<br />

There is now compelling evidence that <strong>charge</strong> separation<br />

from an MLCT excited state into TiO2 can occur faster than<br />

vibrational relaxation. The question of whether this is necessary<br />

or useful has arisen. The injected electrons are expected to<br />

trap at coordinatively unsaturated Ti IV sites <strong>with</strong>in a picosecond.<br />

291–295 If the electron could be collected in an external<br />

circuit prior to trapping, energy that would otherwise be lost<br />

to phonon creation could be harvested. This would allow for<br />

1 sun, AM1.5 light-to-electrical power conversion efficiencies<br />

over the single-junction Shockley–Queisser limit of 31% 296<br />

and up to the ‘hot’-injection-limited efficiency of 66%. 297<br />

Thus, ‘hot’ injection is the first step towards achieving single<br />

junction, solar light-to-electrical power conversion efficiencies<br />

greater than 31%. However, should wave-packet dephasing<br />

108,298 accompany photo-induced electron <strong>transfer</strong> from<br />

the sensitizer to TiO 2 acceptor states, ‘hot’-electron capture<br />

may be impossible. Willig et al. have time-resolved such<br />

thermal relaxation in TiO2 to picoseconds. 243 The dephasing<br />

theory implies that electron <strong>transfer</strong> from the electronic wave<br />

packet in TiO2, representing the initially formed ‘hot’ electron,<br />

to a point contact is virtually impossible. 109,243,244 It is thought<br />

that over time the wave packet expands throughout the film<br />

and that elastic and inelastic scattering events alter the<br />

momentum as well as the electronic and vibrational energy<br />

Fig. 19 A schematic illustrating excitation wavelength-dependent<br />

injection yields for Ru(NH3)5(ina)/TiO2, i.e. 15% <strong>with</strong> 532 nm and<br />

30% <strong>with</strong> 416 nm. Taken from cover artwork of ref. 290.<br />

Fig. 20 A schematic illustrating that excited-state electron injection<br />

and subsequent reduction of a co-bound molecular acceptor, A, was<br />

only thermodynamically possible if injection occurred from the initial,<br />

Franck–Condon excited state and not the thexi state. Formation of A<br />

after pulsed-laser light excitation of Ru(bpy)2(dcbq)/TiO2 was shown<br />

to be due to ‘hot’ electron injection followed by TiO 2-mediated<br />

electron transport to the acceptor. Taken from cover artwork of<br />

ref. 135.<br />

of the electron. Thus, most often this cooled state would reach<br />

the interface for collection or reactivity. The realization of<br />

ultrafast electron <strong>transfer</strong> from/through the semiconductor was<br />

proposed to be feasible only under a specific and unlikely<br />

distribution of ‘hot’ electrons. Therefore, even if the excess<br />

energy is not lost through vibrational relaxation of the excited<br />

state of the sensitizers, it will most likely be lost to phonon<br />

creation on the other side of the interface.<br />

If the TiO2 DOS lie above the excited-state reduction potential<br />

of the sensitizer it could be possible to drive an electron<strong>transfer</strong><br />

reaction that would be thermodynamically unfavorable<br />

from the thexi state, Fig. 20. Such a process was recently<br />

realized <strong>with</strong> heteroleptic sensitizers possessing the previously<br />

mentioned dcbq ligands, such as [Ru(bpy)2(dcbq)] 2+ . 135,299 In<br />

this case, the acceptor was a ground-state sensitizer. The<br />

interfacial energetics were such that Ecb/DOS were more<br />

negative than E o (Ru II/+ ) and E o (Ru III/II* ). Therefore after<br />

‘hot’-electron injection, resulting in a TiO2(e ) and Ru III ,the<br />

TiO2(e ) could recombine <strong>with</strong> the oxidized sensitizer or reduce<br />

another surface-bound sensitizer to form Ru III /TiO2/Ru + .<br />

Using nanosecond transient absorption spectroscopy, spectral<br />

evidence for the allowed processes in Fig. 20 and wavelengthdependent<br />

injection yields lead to the conclusion that ‘hot’<br />

injection was occurring <strong>with</strong> this compound. The conclusion<br />

that formation of Ru III /TiO2/Ru + is mediated by the TiO2<br />

DOS was supported by the fact that there was insufficient free<br />

energy stored in the thexi state of the sensitizer to reduce the<br />

acceptor. Thus, the proposed mechanism was that ‘hot’ injection<br />

was followed by reduction of a TiO2-bound acceptor. This<br />

serves as a proof-of-principle example and suggests that ‘hot’<br />

injection can be used to drive ‘uphill’ redox reactions of<br />

relevance to exceeding the Shockley–Queisser limit.<br />

C Metal-to-particle <strong>charge</strong> <strong>transfer</strong><br />

There is another mechanism of photo-induced electron injection<br />

that has been observed for <strong>metal</strong>-cyano compounds<br />

anchored to TiO 2. This mechanism has been termed <strong>metal</strong>to-particle<br />

<strong>charge</strong> <strong>transfer</strong> (MPCT). This is apparent based on<br />

132 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


the observations that: (a) when said compounds are anchored<br />

to TiO2 a new absorption band is formed that was not present<br />

in fluid solution and (b) light excitation into said absorption<br />

band results in immediate formation of TiO 2(e )/S + . An<br />

interesting feature of such sensitizers is that, by definition,<br />

the injection yield is unity as light absorption and electron<br />

<strong>transfer</strong> to TiO2 are one in the same process. This is in contrast<br />

to injection from MLCT excited states described above whose<br />

injection efficiency has been shown to be a function of the pH,<br />

ionic strength, excitation wavelength, and temperature. 300<br />

MPCT absorption bands were observed for the first time<br />

upon binding M(CN)x n+ complexes to TiO2 nanocrystallites<br />

(M = Fe, Ru, Os, Re, Mo, W). 301,302 Some of these adducts<br />

extended the visible light photoresponse of TiO 2 beyond 700 nm.<br />

Hupp et al. discovered that the resonance Raman spectrum of<br />

Fe(CN) 6/TiO 2 colloids exhibited the coupling of ten vibrational<br />

modes to MPCT, three of which were surface modes. 303,304<br />

Jortner and colleagues have previously described an applicable<br />

theoretical model for describing such multimode electron <strong>transfer</strong>,<br />

305–310 however the coupling of multiple surface modes to<br />

interfacial electron <strong>transfer</strong> was unprecedented experimentally.<br />

Fe II -based coordination compounds containing both MPCT<br />

and MLCT bands of the type [Fe(LL)(CN)4] 2 were studied,<br />

whereLL=bpy,dmb,or4,4 0 -diphenyl-bpy. It was shown that<br />

the MLCT absorption bands were solvatochromic whereas the<br />

MPCT bands were not. 77,311 In light of this and <strong>with</strong> electrochemical<br />

results indicating that E o (Fe III/II ) did shift <strong>with</strong> solvent,<br />

it was suggested that the TiO 2 DOS shifted as well and in a<br />

concerted fashion. There exists a precedence for molecular and<br />

TiO2 reduction potentials shifting in concert when the molecule is<br />

poised <strong>with</strong>in the ionic double layer, i.e. Helmholtz and diffuse<br />

layers. 312,313 Using an [Fe(bpy)(CN)4] 2<br />

sensitizer, the possibility<br />

for a dual-mechanism of sensitization was explored, Fig. 21. 311<br />

Excitation directly into the MPCT band would inherently result<br />

in an injection yield of unity and be independent of the experimental<br />

conditions. In acetonitrile solutions it was shown that the<br />

injection yield could be reversibly tuned <strong>with</strong> the addition of<br />

LiClO4; clearly this was not a result of a MPCT <strong>transition</strong> as the<br />

UV-Vis absorption spectrum was largely independent of electrolyte.<br />

Thus, not only was direct MPCT present in this system, but<br />

less-efficient electron injection from a proximal MLCT excited<br />

state was apparent as well.<br />

Fig. 21 Ball-and-stick models for Fe(bpy)(CN)4/TiO2 portraying two<br />

possible mechanisms for photo-induced electron injection into TiO 2:<br />

(A) direct, <strong>metal</strong>-to-particle <strong>charge</strong> <strong>transfer</strong> (MPCT) sensitization; (B)<br />

sensitization by means of a <strong>metal</strong>-to-ligand <strong>charge</strong>-<strong>transfer</strong> (MLCT)<br />

excitation followed by excited-state electron injection. Taken from<br />

cover artwork of ref. 311.<br />

Intervalence <strong>charge</strong>-<strong>transfer</strong> (IVCT) bands exist for mixedvalence,<br />

polymeric Fe II –CN–Ti IV cyano complexes and are<br />

speculated to be related to MPCT bands in Fe(CN) 6/TiO 2. 314<br />

The MPCT absorption bands were similar in energy and spectral<br />

width to those previously described for outer-sphere <strong>charge</strong><br />

<strong>transfer</strong> <strong>with</strong> iron-cyano anions. From this, a question arises:<br />

does light absorption on TiO2 promote and electron from Fe II to<br />

an adjacent Ti IV site or to a Ti IV site <strong>with</strong>in the interior of a<br />

nanocrystallite? By electroabsorption (Stark) spectroscopy it was<br />

determined that the MPCT distance was 5.3 A ˚ basedonthe<br />

dipole moment change and the Liptay treatment. 314 This was<br />

<strong>with</strong>in error of the distance from Fe to Ti using molecular<br />

modeling on [(CN)5Fe II –CN–Ti IV (H2O)4O] 2 , although it was<br />

slightly larger than empirical values measured for related<br />

Fe II –CN–M compounds. Similar distances were found for<br />

Mo-, Ru- and W-cyano complexes on TiO 2 and all support the<br />

hypothesis that MPCT bands represent electronic <strong>transition</strong>s to<br />

an orbital on a Ti atom that is in proximity to the bound cyano<br />

nitrogen atom. 315 Additionally, based on the above calculated<br />

distance and the fact that the free NC ligands are even further<br />

from the surface than the <strong>metal</strong> center, identification of the<br />

process as MPCT and not ligand-to-<strong>metal</strong> <strong>charge</strong> <strong>transfer</strong><br />

(LMCT), i.e. from NC to Ti, was substantiated.<br />

Some organic bases are also known to display direct, <strong>charge</strong><strong>transfer</strong><br />

absorption bands when anchored to TiO 2, the most<br />

well-known being catechol. 188,300 Two Os II -polypyridyl compounds<br />

<strong>with</strong> bpy-catechol derivatives for surface attachment<br />

were recently reported. 316 Absorption features assigned as<br />

direct catechol-to-particle <strong>charge</strong> <strong>transfer</strong> and MLCT on TiO2<br />

and ZrO2 (a wide-bandgap semiconductor <strong>with</strong> EBG = 4.5 eV)<br />

were observed and PL assigned to radiative <strong>charge</strong> recombination<br />

was reported. Similar PL behavior was previously reported<br />

for organic compounds anchored to TiO2, but the assignment<br />

of this to radiative recombination was later questioned. 300,317<br />

D Reduced-sensitizer injection<br />

An alternative mechanism exists for photo-induced electron<br />

injection wherein the excited-state is reduced prior to interfacial<br />

<strong>charge</strong> separation, Fig. 22(b). This results in injection<br />

from a non-electronically excited sensitizer. For this reason<br />

DSSCs operating under this mechanism would appropriately<br />

be termed regenerative galvanic cells as injection would be a<br />

dark, thermodynamically favorable process. 318 This alternative<br />

sensitization method has been called supersensitization; 153<br />

the donor is termed the supersensitizer due to its requirement<br />

in achieving effective overall sensitization. 319 A unique aspect<br />

of this mechanism is that the oxidized form of the sensitizer is<br />

never generated. Therefore, it may be particularly well-suited<br />

for sensitizers that are unstable in their oxidized forms. An<br />

advantage <strong>with</strong> MLCT excited states is that the reduced form<br />

of the sensitizer is a stronger reductant than the MLCT excited<br />

state, typically by 200 to 400 mV.<br />

Kirsch-De Mesmaeker and co-workers first reported compelling<br />

evidence for reduced ruthenium sensitizers <strong>transfer</strong>ring<br />

electrons to SnO2 electrodes. 319 The coincidence of Stern–<br />

Volmer constants measured by analysis of the photocurrent<br />

enhancement and PL quenching <strong>with</strong> hydroquinone donors left<br />

little doubt as to the sensitization mechanism. Additional<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 133


Fig. 22 Ball-and-stick models for Ru(bpy)2(dcb)/TiO2 depicting the<br />

two possible mechanisms for photo-induced electron injection into<br />

TiO2 and regeneration of the sensitizer: (A) MLCT excited-state<br />

electron injection followed by regeneration of the oxidized sensitizer<br />

by the solution donor, D; (B) Reduced sensitizer injection that results<br />

from reductive quenching of the excited state by D followed by dark<br />

electron injection from the reduced sensitizer. Taken from Scheme 1 of<br />

ref. 320.<br />

spectroscopic evidence for photo-induced electron injection by<br />

reduced sensitizers was reported for Ru(bpy)2(dcb)/TiO2 in<br />

0.1 M TBAClO4 acetonitrile electrolyte containing neutral,<br />

organic phenothiazine (PTZ) electron donors. 321 Nanosecond<br />

transient absorption data demonstrated the rapid formation of<br />

TiO 2(e )/PTZ + , while in the absence of PTZ there was littleto-no<br />

evidence for injection. Injection was rate limited by<br />

diffusional quenching of the MLCT excited state so the<br />

Ru II (bpy) 2(dcb )/TiO 2 intermediate was not directly observed.<br />

An interesting case of photo-induced, reduced-sensitizer<br />

electron injection was reported <strong>with</strong> the bi<strong>metal</strong>lic sensitizer<br />

(bpy)2 Ru II –BL–Rh III (dcb)2/TiO2, BL = 1,2-bis[4-(4 0 -methylbpy)]ethane,<br />

Fig. 23(a). 322 About two-thirds of the MLCT<br />

excited states of the ruthenium chromophore were quenched<br />

by electron <strong>transfer</strong> to the Rh(dcb)2 group to form a<br />

(bpy)2Ru III –BL–Rh II (dcb)2/TiO2 <strong>charge</strong>-separated state while<br />

the remaining directly injected an electron into TiO2.** This<br />

observed branching ratio was proposed to result from different<br />

surface orientations. Approximately 40% of the intramolecular,<br />

<strong>charge</strong>-separated state, (bpy) 2Ru III –BL–Rh II (dcb) 2/TiO 2,<br />

injected electrons into TiO2 to form (bpy)2Ru III –BL–<br />

Rh III (dcb)2/TiO2(e ), while the remaining underwent backelectron<br />

<strong>transfer</strong> to form ground-state products, Fig. 23(b).<br />

The Ru II* -based injection occurred <strong>with</strong>in the time resolution<br />

of the instrument, i.e. o10 ns, while the Rh II -based injection<br />

occurred in o100 ns following light excitation.<br />

In order to realize efficient DSSCs that operate by this<br />

mechanism, sensitizers that are potent photo-oxidants must<br />

be utilized. This stems from that fact that the I 3 /I redox<br />

mediator is the only redox mediator that yields high light-toelectrical<br />

power conversion efficiencies and the I /I reduction<br />

potential is rather positive. Ru II sensitizers that are strong<br />

excited-state oxidants can be prepared <strong>with</strong> ligands such as<br />

2,2 0 -bipyrazine (bpz). For example, the E o (Ru 2+*/+ ) of<br />

[Ru(bpz)2(deeb)] 2+ was found to be greater than +1.0 V vs.<br />

SCE 69,70,323 (+1.24 V vs. NHE 229 ). While the excited state of<br />

this and related sensitizers were found to be efficiently<br />

quenched by iodide or phenothiazine donors, the reduced form<br />

** We emphasize that while the scheme and this abbreviation imply<br />

that the reduction is <strong>metal</strong> based, it may in fact be ligand localized, i.e.<br />

on a dcb.<br />

Fig. 23 (A) A diagram of a [(bpy) 2Ru II –BL–Rh III (dcb) 2] 5+ sensitizer<br />

bound to a TiO2 nanocrystallite. (B) A schematic depicting the relative<br />

redox energies, lifetimes, and quantum yields for each step in the<br />

photo-induced injection process. Taken from Fig. 10 and 12, respectively,<br />

of ref. 322.<br />

of the compound that resulted, Ru II (bpz)(bpz )(deeb)/TiO 2,<br />

did not inject electrons into TiO 2. 320 In fact, very similar<br />

transient absorption features were observed in solution,<br />

on TiO2, and on ZrO2, while extremely small photocurrents<br />

(IPCE o 10 4 ) were observed in DSSCs. Some improvement<br />

was observed when the semiconductor was changed to SnO2,<br />

but the injection yields remained poor. 323<br />

Another interesting case of reductive quenching of an<br />

excited state that did not result in electron injection was<br />

reported for the mono-anion of Z907/TiO 2 in the presence<br />

of a high concentration of 1-propyl-3-methylimidazolium<br />

iodide. 324 A new transient spectroscopic feature was discovered<br />

that was attributed to the reduced sensitizer, which<br />

presumably formed by reductive quenching of the excited state<br />

by iodide. The decay of this species was attributed to a backreaction<br />

<strong>with</strong> I3 (t1/2 = B1 ms) yet it is not clear why this<br />

state did not inject electrons into TiO2.<br />

4. Sensitizer regeneration<br />

A Intramolecular regeneration<br />

Considerable effort has been set forth to regenerate the<br />

oxidized sensitizer by intramolecular electron <strong>transfer</strong>. This<br />

could be considered a ‘‘hole’’ <strong>transfer</strong> reaction that translates<br />

the oxidizing equivalent away from the Ru III -<strong>metal</strong> center and,<br />

ideally, the TiO2 surface. Very similar mechanisms are wellknown<br />

in the field of supramolecular photochemistry where<br />

compounds of the type (D)n–C–(A)m are often employed ((D)n<br />

are donor molecules, C is a chromophore/sensitizer, and (A)m<br />

are acceptor molecules). When solely two components are<br />

present the compounds are termed dyads. 88,94,325,326 At TiO 2<br />

interfaces a variety of C*–D dyads have been characterized;<br />

to our knowledge, Wrighton and co-workers were the first to<br />

134 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


attach a dyad to a semiconductor electrode. 327 The ability to<br />

control hole-<strong>transfer</strong> reactions at the molecular level is important<br />

for many classes of solar cells. One can envision<br />

future-generation DSSCs where multiple hole-<strong>transfer</strong> steps<br />

translate the oxidizing equivalent from the sensitized interface<br />

directly to a counter electrode thereby eliminating the need for<br />

the solution-based redox mediators that are required today,<br />

e.g. I3 /I .<br />

In practice there are at least two ways in which D–C*/TiO2<br />

- D + –C/TiO2(e ) reactions can occur. They correspond to<br />

<strong>charge</strong>-separation mechanisms from the excited or reduced<br />

states that were described in sections 3/B and 3/D, respectively.<br />

Since C* is a weaker oxidant than C + ,itispossibletodesign<br />

dyads covalently bound to weak donors that only react by the<br />

first mechanism. When strong electron donors are used, the<br />

mechanistic pathway is dependent on the relative rate constants<br />

for excited-state electron injection and intramolecular <strong>charge</strong><br />

separation. Since excited-state injection is often found to be<br />

ultrafast, the first mechanism probably predominates even<br />

though it cannot always be unambiguously identified.<br />

It should be pointed out that in some regards N3/TiO2 is<br />

thought to undergo a similar intramolecular, <strong>charge</strong>-<strong>transfer</strong><br />

process. DFT calculations for N3 + predict considerable hole<br />

density on the isothiocyanate ligands. 56–58 It is also known<br />

from electrochemical measurements that there are two closely<br />

spaced oxidations for N3, the first is predominantly <strong>metal</strong><br />

based while the second is mainly isothiocyanate based. Therefore,<br />

in the <strong>charge</strong>-separated state, N3 + /TiO 2(e ), there is<br />

likely some partial ‘‘hole <strong>transfer</strong>’’ from the Ru III -<strong>metal</strong> center<br />

to the isothiocyanate ligands. In most of the examples discussed<br />

below, the electronic coupling between the electron<br />

donor and the Ru-<strong>metal</strong> center is much weaker, giving rise to<br />

complete hole hopping rather than partial <strong>charge</strong> <strong>transfer</strong>.<br />

i Organic donors. In collaboration <strong>with</strong> Bignozzi and his<br />

research group in Ferrara, Italy, we reported the first timeresolved<br />

spectroscopic studies of intramolecular sensitizer<br />

regeneration <strong>with</strong> the dyad [Ru(4-CH3,4 0 -CH2-PTZbpy)(dcb)2]<br />

2+ . 328,329 Comparative studies in fluid methanol<br />

solution, visible-light excitation of this dyad resulted in the<br />

creation of the MLCT excited state that was quickly quenched<br />

by electron <strong>transfer</strong> from the PTZ group. The reductive<br />

excited-state quenching was moderately exergonic (o0.25 eV)<br />

and had an approximate rate constant of B2.5 10 8 s 1 in<br />

methanol. The corresponding <strong>charge</strong>-recombination step was<br />

faster than the quenching by PTZ and thus there was little<br />

appreciable spectroscopic observation of the electron-<strong>transfer</strong><br />

product.<br />

When the dyad was anchored to TiO 2 thin films and<br />

immersed in acetonitrile, MLCT excitation resulted in a new<br />

<strong>charge</strong>-separated state <strong>with</strong> an electron in TiO2 and an oxidized<br />

PTZ group, abbreviated PTZ + -Ru II /TiO2(e ). It was<br />

not possible to determine the mechanism of <strong>charge</strong> separation<br />

yet the authors speculated that after excited-state electron<br />

injection, electron <strong>transfer</strong> from PTZ to the Ru III -<strong>metal</strong> center<br />

( DG B 0.36 eV) produced PTZ + -Ru II /TiO2(e ). Recombination<br />

of TiO 2(e )s <strong>with</strong> PTZ + to yield ground-state products<br />

occurred <strong>with</strong> a rate constant of 3.6 10 3 s 1 . Excitation of a<br />

model compound that did not contain the PTZ donor under<br />

otherwise identical conditions gave rise to the immediate<br />

formation of a <strong>charge</strong>-separated state, Ru III /TiO2(e ), whose<br />

recombination kinetics were complex and analyzed by a distribution<br />

model <strong>with</strong> an average rate constant of 3.9 10 6 s 1 .<br />

Therefore, translating the ‘‘hole’’ from the Ru III -<strong>metal</strong> center<br />

to the pendant PTZ moiety slowed TiO 2(e ) recombination by<br />

about three orders of magnitude. This work provided an<br />

example of how the principles of stepwise <strong>charge</strong> separation,<br />

originally developed in the field of supramolecular photochemistry,<br />

can be applied to solid-state materials.<br />

Shortly thereafter, Grätzel and co-workers reported dyads that<br />

could undergo intramolecular ‘‘hole’’ <strong>transfer</strong> after excited-state<br />

electron injection. 330 These authors emphasized the significant<br />

color changes that accompanied electron <strong>transfer</strong> and potential<br />

applications in photochromic devices. Interestingly, they observed<br />

long-lived <strong>charge</strong>-separation, like the PTZ + -Ru II /<br />

TiO 2(e ) described above, in some cases while not in others.<br />

Since that time a number of dyads have been attached to<br />

TiO2 and are discussed further below. A commonly utilized<br />

electron donor is a triarylamine moiety, NAr3. Three<br />

Ru II -NAr3-type sensitizers <strong>with</strong> tpy-based ligands were<br />

studied in order to determine the optimal spatial location of<br />

the amine donor in relation to the excited-state electron. 330<br />

One compound had a 4-(N,N-di-p-anisylamino)phenyl group<br />

conjugated to a second tpy ligand, the second contained a<br />

benzyl ether interlocking group between the same amine donor<br />

and the tpy ligand, and the third compound contained only a<br />

4,4 0 ,4 00 -trimethyl-tpy ligand, Fig. 24. The latter compound was<br />

also co-adsorbed <strong>with</strong> a donor moiety bound to a phosphonated<br />

ether. Using long-wavelength resonance Raman spectroscopy<br />

it was deduced that the excited electron in the excited<br />

state of the first compound was located on the donor-containing<br />

tpy ligand whereas for the other two compounds it was located<br />

on the surface-bound phosphonated-tpy ligand. For all three<br />

compounds, photo-induced electron injection occurred<br />

quantitatively in o1 ns in air. In propylene carbonate, the<br />

quantum yield for formation of NAr 3 + -Ru II /TiO2(e ) was<br />

only 0.60 and occurred <strong>with</strong>in 20 ns for the first compound<br />

while for the second compound it was unity <strong>with</strong> biphasic<br />

kinetics, a 10 ns component and a 100 ns component.<br />

The third compound <strong>with</strong> the co-bound donor experienced<br />

practically unity conversion to the NAr3 + /TiO2(e ) <strong>charge</strong>separated<br />

state. This study illustrated that remote injection<br />

was less efficient than injection from a surface-bound ligand.<br />

Some remarkably long-lived <strong>charge</strong>-separated states<br />

were observed after pulsed-light excitation of similar compounds.<br />

By covalently attaching the ether-NAr 3 donor<br />

group just described above to a dmb ligand in a<br />

cis-Ru(dmb-X)(dcb)(NCS) 2/TiO 2 system, the half-life of the<br />

<strong>charge</strong>-separated state was found to be over half a second, as<br />

shown schematically, Fig. 25(a). 331 Haque, Durrant, and<br />

colleagues increased this lifetime further by employing<br />

Ru(4,4 0 -(R)2-bpy)(dcb)2/TiO2 systems, where R contained<br />

one triphenylamine group (NPh3), two NPh3 groups, or a<br />

poly(vinyl-NPh3) group of about 100 units, Fig. 25(b). 332 The<br />

introduction of about 100 amines increased the half-life of<br />

the <strong>charge</strong>-separated state to over 4 seconds, as compared to<br />

350 ms and 5 ms for the other two compounds, respectively.<br />

The kinetics for excited-state electron injection and subsequent<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 135


Fig. 24 Chemical structures of sensitizers containing intramolecular,<br />

or nearby, organic donors so as to increase the <strong>charge</strong>-separation<br />

distance. Taken from Fig. 5 of ref. 330.<br />

hole <strong>transfer</strong> from the Ru III -<strong>metal</strong> center to the covalently<br />

bound NPh3 moiety occurred <strong>with</strong>in the instrument response<br />

time, i.e. B10 ns.<br />

N3 derivatives, cis-Ru(4,4 0 -(R)2-bpy)(dcb)(NCS)2, R =<br />

NPh3 or CH3, bound to TiO2 thin films were examined in order<br />

to study the effects of Ru III -hole <strong>transfer</strong> to a triphenylamine<br />

moiety. 82 Unexpectedly, both sensitizers exhibited similar transient<br />

features providing no evidence for hole <strong>transfer</strong> in the<br />

former sensitizer. Not<strong>with</strong>standing, the photoelectrochemical<br />

properties of the two sensitized thin-film electrodes differed<br />

Fig. 25 (A) A schematic depicting the cis-Ru(dmb-ether-NAr 3)-<br />

(dcb)(NCS)2 sensitizer bound to a TiO2 nanocrystallite and the overall<br />

mechanism for photoinduced <strong>charge</strong> separation and recombination<br />

<strong>with</strong> corresponding time scales. Taken from cover artwork of ref. 331.<br />

(B) The chemical structure of the sensitizer employed to increase<br />

the half-time of the S + /TiO2(e ) <strong>charge</strong>-separated state to over 4 s<br />

(n = 100). Taken from Scheme 1 of ref. 332.<br />

Fig. 26 The chemical structure of the sensitizer employed to study<br />

intramolecular <strong>charge</strong> separation on TiO2 thin films. The hole was<br />

successfully <strong>transfer</strong>red away from the Ru III -<strong>metal</strong> center and TiO 2<br />

surface to the carotenoid moiety. Taken from Scheme 3 of ref. 333.<br />

significantly and a much larger Voc was measured for the<br />

NPh 3-containing sensitizer. The authors speculated that the<br />

enhanced V oc resulted from a larger dipole that was nascently<br />

formed on the sensitizer bearing the NPh 3 moiety. Based on the<br />

enhanced extinction coefficient of this dye and the results<br />

obtained when the small-perturbation Voc-decay technique<br />

was employed, it was proposed that photo-induced electron<br />

injection into the TiO2 acceptor states and partial hole delocalization<br />

from the Ru III -<strong>metal</strong> center to the NPh3 moiety<br />

occurred in one concerted step. Thus, increased <strong>charge</strong> separation<br />

could be achieved concomitant <strong>with</strong> electron injection by<br />

partial delocalization of the hole on the ligand.<br />

A series of novel sensitizers each containing a 4 0 -X-tpy<br />

ligand (X = Ph–PO 3(C 2H 5) 2, Ph–PO 3H 2, PO 3(C 2H 5) 2,<br />

PO 3H 2, or COOH) bound to TiO 2 were studied <strong>with</strong> hopes<br />

of increasing the <strong>charge</strong>-separation distance between the<br />

TiO2(e ) and oxidized sensitizer. 333 Only the sensitizer shown<br />

in Fig. 26, containing a phenyl-amide-carotenoid bound to a<br />

second tpy, provided unequivocal evidence for intramolecular<br />

sensitizer regeneration and thus increased <strong>charge</strong> separation,<br />

which was complete in o10 ns. However, even though DFT<br />

calculations indicated that the LUMO and the MLCT excited<br />

state were located on the 4 0 -phenylphosphonate-tpy ligand<br />

which was bound to TiO2, injection yields were poor. It was<br />

postulated that the out-of-plane phenyl spacer gave poor<br />

electronic coupling to TiO 2.<br />

The compound [Ru(BTL)(deeb) 2] 2+ , where BTL is 9 0 -[4,5bis(cyanoethylthio)]-1,3-dithiol-2-ylidene]-4<br />

0 ,5 0 -diazafluorene,<br />

was found to have an extinction coefficient almost three times<br />

as large as Ru(bpy)3 2+ in the visible region. 81 Interestingly, the<br />

transient absorption features in solution and on TiO2 differed<br />

greatly. In solution, a transient state was observed <strong>with</strong><br />

spectroscopic properties characteristic of an MLCT excited<br />

state, <strong>with</strong> t =25nsat 40 1C, whereas when bound to TiO2<br />

a large positive absorption feature near 520 nm was observed<br />

and assigned to the oxidized dithiolene ligand. In fluid solution<br />

the driving force for reductive quenching of the MLCT excited<br />

state was unfavorable. However, when anchored to TiO 2,an<br />

electron was injected and the hole had translated from the<br />

Ru III -<strong>metal</strong> center to the dithiolene-containing ligand, <strong>with</strong>in<br />

10 ns after light excitation, Fig. 27.<br />

ii Transition-<strong>metal</strong> donors. Although an organic donor is<br />

more optimal for practical applications, an advantage of using a<br />

<strong>transition</strong> <strong>metal</strong> as the donor is that its redox potential can be<br />

more easily tuned over wide energies by utilizing different ligands.<br />

The bi<strong>metal</strong>lic sensitizer [Cl(bpy)2Os II –bpa–Ru II (dcb)2Cl] 2+ ,<br />

abbreviated Os-bpa-Ru, where bpa is 1,2-bis(4-pyridyl)ethane,<br />

was anchored to TiO 2. 334 Pulsed 532 nm or 416 nm light<br />

excitation of a Os-bpa-Ru/TiO 2 thinfilmimmersedin1.0M<br />

136 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 27 A schematic depicting a novel, high extinction coefficient<br />

sensitizer bound to a TiO 2 nanocrystallite and photo-induced electronand<br />

hole-<strong>transfer</strong> mechanisms. This sensitizer is unique in that the<br />

extended conjugation on the dithiolene-containing ligand is in the 3<br />

and 3 0 positions. Taken from cover artwork of ref. 81.<br />

LiClO4 acetonitrile electrolyte resulted in rapid excited-state<br />

electron injection (Ru II* +TiO2 - Ru III +TiO2(e )) and<br />

intramolecular electron <strong>transfer</strong> (Os II –bpa–Ru III - Os III –<br />

bpa–Ru II ) to ultimately form an interfacial <strong>charge</strong>-separated<br />

state <strong>with</strong> a TiO 2(e ) and an oxidized Os III -<strong>metal</strong> center, Os III –<br />

bpa–Ru/TiO 2(e ). This same state was also generated after<br />

selective 3 MLCT excitation of the Os II moiety <strong>with</strong> 683 nm light.<br />

The rates of intramolecular and interfacial electron <strong>transfer</strong> were<br />

fast, k 4 10 8 s 1 , while interfacial <strong>charge</strong> recombination,<br />

Os III –bpa–Ru/TiO2(e ) - Os II –bpa–Ru/TiO2, required milliseconds<br />

for completion. The results show a general strategy for<br />

promoting rapid intramolecular hole <strong>transfer</strong> (Os II –bpa–Ru III -<br />

Os III –bpa–Ru II ) after excited-state electron injection and a ‘remote,’<br />

excited-state electron-injection process that occurs after<br />

direct excitation of the Os II chromophore, whose thexi state<br />

possesses far too little energy to <strong>transfer</strong> energy to the ruthenium<br />

moiety.<br />

Related studies <strong>with</strong> (bpy)2M II –bpt–Ru II (dcb)2/TiO2 thin<br />

films (M = Ru or Os), abbreviated M–bpt–Ru, where bpt-H<br />

= 3,5-bis(pyridin-2-yl)-1,2,4-triazole, showed evidence for<br />

two different electron-injection mechanisms depending on<br />

M. 335 For the all ruthenium compound, excited-state energy<br />

<strong>transfer</strong> to the TiO2-bound, dcb-containing, ruthenium moiety<br />

followed by excited-state injection was deduced based on<br />

transient PL and absorbance measurements. Although not<br />

directly observed, hole <strong>transfer</strong> to the proximal ruthenium<br />

moiety was thermodynamically favorable after excited-state<br />

injection. For the M = Os compound, excitation into the<br />

Ru II -based MLCT band resulted in excited-state energy <strong>transfer</strong><br />

to the proximal osmium moiety prior to injection, and after<br />

remote injection the hole was proposed to remain on the<br />

Os-<strong>metal</strong> center. The expected Os III –bpt–Ru II /TiO 2(e )<br />

product formed <strong>with</strong>in the laser pulse (B10 ns). It was<br />

proposed that since some of the exciting light was absorbed<br />

by the Os II moiety, and energy <strong>transfer</strong> to the ruthenium<br />

moiety was energetically unfavorable, some remote injection<br />

from the Os-localized excited state also occurred in this case.<br />

Studies <strong>with</strong> a solution and surface-bound trinuclear ruthenium<br />

complex, (Ru III –Ru II )(L)–amide–(bpy)Ru II (dcb)2/TiO2,<br />

revealed that MLCT excitation of the mononuclear Ru II -<strong>metal</strong><br />

center resulted in a transient absorption spectrum indicative of<br />

(Ru III –Ru II )(L)–amide–(bpy)Ru III (dcb) 2/TiO 2(e ), Fig. 28. 336<br />

This intramolecular <strong>charge</strong>-separated compound was completely<br />

formed by 200 ps, at which time the injection yield was<br />

deemed to be o10%. However, by 300 ns a spectrum consistent<br />

<strong>with</strong> (Ru III –Ru III )(L)–amide–(bpy)Ru II (dcb)2/TiO2(e )<br />

was observed and was shown to have a half-life, t1/2 = B1 ms.<br />

This illustrates that slow hole <strong>transfer</strong> can occur over large<br />

distances under the appropriate conditions.<br />

Coordination compounds of the form [(LL)(L 0 L 0 )Ru II -<br />

(BL 0 )Ru II (LL)(L 0 L 0 )] n+ (n = 2, 3 depending on the number<br />

of deprotonated carboxylic acid functional groups) were investigated<br />

on TiO 2, where LL and L 0 L 0 are bpy and/or dcb<br />

and BL 0 is a bridging ligand: either tetrapyrido[3,2-a:2 0 ,3 0 -<br />

c:3 00 ,2 00 -h:2 000 ,3 000 -j]phenazine (tpphz) or 1,4-bis(phen-[5,6-d]imidazol-2-yl)benzene<br />

(bfimbz), where phen is 1,10-phenanthroline.<br />

337 As the BL 0 ligands are rigid and linear heteroaromatic<br />

entities, remote, excited-state electron injection could<br />

be examined <strong>with</strong> little fear of unexpected outer-sphere<br />

ligand–surface interactions due to ligand flexibility. It was<br />

shown that when BL 0 was tpphz—a ligand possessing p*<br />

energetics below the p* levels of the surface-bound dcb<br />

ligand—injection could be time resolved due to the thexi state<br />

being localized on tpphz, away from a surface-bound dcb<br />

ligand and <strong>with</strong> less reducing power for injection. However,<br />

this slow injection was found to be not only distance- and/or<br />

driving force-dependent but orientation-dependent as well.<br />

When [(bpy)(dcb)Ru II (tpphz)Ru II (bpy)(dcb)] n+ (n = 2 or 3)<br />

was employed as the sensitizer injection could be time-resolved<br />

Fig. 28 A schematic depicting a sensitizer employed to study intramolecular<br />

<strong>charge</strong> separation on TiO2 thin films. Interestingly, slow<br />

intramolecular <strong>charge</strong> separation between the mononuclear Ru III and<br />

dinuclear Ru II –Ru III could be observed on the hundreds of nanoseconds<br />

time scale. Taken from cover artwork of ref. 336.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 137


Fig. 29 Density Functional Theory (DFT) optimized geometry for<br />

two bi<strong>metal</strong>lic Ru II compounds. When a distal ligand possessed<br />

carboxylic acid functional groups capable of binding to the TiO2<br />

surface, the geometry of the minimized energy configuration had it<br />

binding to the surface as well (b). Although the LUMO of the doubly<br />

bound form was spatially closer to the TiO2 surface, the singly bound<br />

sensitizer had a faster injection rate due to better electronic coupling<br />

<strong>with</strong> the TiO2 DOS. Taken from Fig. 7 of ref. 337.<br />

using nanosecond transient absorption spectroscopy, whereas<br />

<strong>with</strong> [ (bpy)2Ru II (tpphz)Ru II (bpy)(dcb)] n+ (n = 2 or 3) it<br />

could not, kinj 410 8 s 1 . Using DFT geometry optimization<br />

software it was hypothesized that electronic coupling, and not<br />

distance from the TiO 2 surface, could explain the differences,<br />

Fig. 29. The location of the p* orbital of the heterobinuclear<br />

complex in relation to the TiO 2 surface allowed for better<br />

electronic coupling between the sensitizer and the TiO2 DOS<br />

even though the Nphenazine–Ti distance was increased by over a<br />

factor of two. Photoelectrochemical measurements supported<br />

this and indicated that by increasing the distance for backelectron<br />

<strong>transfer</strong> the photocurrent efficiency could be enhanced.<br />

B Intermolecular regeneration<br />

In DSSCs, redox mediators are added to the external electrolyte.<br />

The reduced form of the mediator must regenerate the<br />

oxidized sensitizer by electron <strong>transfer</strong> prior to recombination<br />

<strong>with</strong> the injected electron. The oxidized form of the redox<br />

mediator is then reduced at the platinum counter electrode, a<br />

process not described herein. Ideally, all redox states of the<br />

redox mediator would not competitively absorb light.<br />

Although ion-pairing or surface adsorption <strong>with</strong> such mediators<br />

may occur, for the organization of this review we consider<br />

these to be intermolecular electron-<strong>transfer</strong> reactions.<br />

i Regeneration by iodide<br />

a Sensitizers in solution. By far the most effective donor in<br />

DSSCs is iodide. 12 All confirmed reports of light-to-electrical<br />

power conversion efficiencies over 10% utilize iodide and<br />

state-of-the-art DSSCs require iodide. 12 While many of the<br />

details of iodide oxidation at sensitized electrodes are now<br />

becoming available, it is important to point out that the<br />

aqueous redox chemistry of iodide and homogeneous reactions<br />

<strong>with</strong> <strong>transition</strong>-<strong>metal</strong> compounds have long been<br />

known. 338–342<br />

Shown in Scheme 1 is a Latimer-type diagram for the<br />

aqueous redox chemistry of iodide. Additional values and<br />

details are available in the review by Stanbury. 338 The formal<br />

one-electron reduction potential of the iodine atom is very<br />

positive, E o (I /I ) = +1.33 V vs. NHE. 338 Therefore, a potent<br />

oxidant is required to generate iodine atoms. However, another<br />

pathway exists in which two iodides can be oxidized<br />

directly to I2 , E o (I2 ) = +1.03 V vs. NHE. 338 Based on<br />

potentials alone, it is tempting to conclude that this latter<br />

pathway is the only mechanism available to oxidized sensitizers<br />

like N3 + , since generation of iodine atoms would be<br />

thermodynamically unfavorable by close to 250 mV. 25 However,<br />

it should be kept in mind that the potentials listed are for<br />

standard-state conditions in aqueous electrolytes and that<br />

adsorption to the TiO 2 surface may have a significant effect.<br />

Walter and Elliott have provided evidence that interactions<br />

between iodide and the bpy ring may also activate iodide. 343<br />

Furthermore, the values given in Scheme 1 are for aqueous<br />

solutions. Since there is good reason to believe that the<br />

reduction potentials will vary significantly <strong>with</strong> solvent, it<br />

would be tremendously helpful to this field if a corresponding<br />

Latimer-type diagram in acetonitrile was available since there<br />

is good reason to believe that the reduction potentials will vary<br />

significantly <strong>with</strong> solvent. For example, the equilibrium constant<br />

for reaction (9) is reported to be 410 6 M 1 in<br />

CH 3CN 344–349 but is only 700–800 M 1 in water. 350<br />

I +I2 " I3 (9)<br />

It is not trivial to obtain the one-electron reduction potentials<br />

experimentally. We and others before us have found that only<br />

two-electron redox processes are observed by voltammetry<br />

measurements at <strong>metal</strong> electrodes. 344,351,352 Stanbury has examined<br />

iodide oxidation by a series of Fe III compounds in<br />

acetonitrile and from this, the sole one-electron <strong>transfer</strong><br />

reduction potential available in acetonitrile that we are aware<br />

of was established through kinetic inhibition measurements,<br />

E o (I /I ) = +0.60 0.01 V vs. the ferrocenium/ferrocene<br />

redox couple (FeCp2 +/0 ) 353 (+1.15 vs. NHE 229 ).<br />

Scheme 1<br />

138 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


The <strong>transition</strong>-<strong>metal</strong> redox chemistry of iodide has previously<br />

been reviewed. 341,342 Two mechanisms have been<br />

observed, based on reactions (10) and (11):<br />

Mox +I - Mred +I (10)<br />

Mox +2I - Mred +I2 (11)<br />

Both are first order in <strong>transition</strong>-<strong>metal</strong> compound, M ox, while<br />

(10) is first order in iodide and (11) is second order in iodide.<br />

Proposed mechanisms for (11), the overall third-order reaction,<br />

include I reacting <strong>with</strong> an [M ox,I ] ion-pair or M ox <strong>with</strong><br />

an [I ,I ] ion-pair. A wide variety of <strong>transition</strong>-<strong>metal</strong> compounds<br />

have been studied and linear free-energy relations for<br />

both reactions now exist. In some cases, <strong>with</strong> mild oxidants<br />

such as Mox = Os(bpy)3 3+ , the reverse reactions became<br />

significant. 339–341<br />

Much less is known about MLCT excited-state oxidation of<br />

iodide. The alternative, reduced-sensitizer electron-injection<br />

process requires interactions of the excited state and iodide.<br />

Early studies <strong>with</strong> [Ru III (bpy) 2(bpy )] 2+ * revealed very inefficient<br />

electron <strong>transfer</strong>, i.e. 1 10 6 M 1 s 1 . 354,355 Interestingly,<br />

excited-state quenching of [Ru III (bpy) 2(dcb )] 2+ *<br />

anchored to SiO2 appears to be somewhat more efficient, i.e.<br />

1 10 8 M 1 s 1 . 356<br />

We recently found that excited-state electron-<strong>transfer</strong> reactions<br />

<strong>with</strong> iodide were significant when ion-paired <strong>with</strong><br />

the ground-state sensitizer. 133,357 Addition of iodide to a<br />

dichloromethane solution of [Ru(bpy)2(deeb)] 2+ resulted in<br />

significant changes to the ground-state absorption spectrum.<br />

A decrease in PL and excited-state lifetime accompanied the<br />

absorption changes consistent <strong>with</strong> both static- and dynamicquenching<br />

mechanisms, respectively. A Benesi–Hildebrandtype<br />

analysis of these absorption changes yielded equilibrium<br />

constants for ion-pairing that were <strong>with</strong>in experimental error<br />

the same as those abstracted from PL quenching data,<br />

Keq = 59 700 M 1 . Similar behavior was observed in<br />

acetonitrile and/or <strong>with</strong> Ru(bpy)3 2+ , however an iodide<br />

concentration that was two orders of magnitude larger was<br />

required. Transient absorption measurements clearly showed<br />

an electron-<strong>transfer</strong> mechanism <strong>with</strong> the appearance of I2<br />

and no evidence for intermediate iodine atom formation; thus<br />

the mechanism appeared to follow reaction (11). The cage<br />

escape yields were low, f = 0.25, but increased to 0.50 <strong>with</strong><br />

Ru(bpy) 3 2+ . Remarkably, the solid-state crystal structure of<br />

Ru(bpy) 2(deeb)I 2 had both iodides associated <strong>with</strong> the carbonyl<br />

oxygens of the ester groups, Fig. 30. One might have<br />

anticipated that Coulombic repulsion would have resulted in a<br />

larger inter-ionic distance then the B6 A ˚ observed. If a similar<br />

structure exists in solution the iodides would be well-positioned<br />

for a concerted reduction of [Ru III (bpy)2(deeb )] 2+ * and<br />

formation of I2 . This is an intriguing possibility as excitedstate<br />

reactions that form chemical bonds are rare in all of<br />

photochemistry. Although, evidence for intermediate I formation<br />

by reaction (10) has recently been observed in our labs. 358<br />

b Sensitizer/TiO2 systems. The first <strong>heterogeneous</strong> reduction<br />

of Ru III -polypyridyl compounds by iodide was reported<br />

by Fitzmaurice and Frei. 359 Photo-induced electron injection<br />

into colloidal TiO 2 from [Ru III (dcb) 2(dcb )] 2+ * was followed<br />

Fig. 30 Space-filling representation of the crystal structure of a single<br />

sensitizer determined by X-ray diffraction showing two iodides associated<br />

<strong>with</strong> the deeb ligand in [Ru(bpy) 2(deeb)] 2+ . This geometry<br />

would allow for facile reductive quenching of the excited or oxidized<br />

forms of the molecule and the proximity of a second iodide could favor<br />

I2 generation as per eqn (11). Taken from cover artwork of ref. 133.<br />

by oxidation of iodide in acidic aqueous solution. From the<br />

pseudo-first-order transient kinetics in 0.5 to 100 mM KI,<br />

a second-order rate constant for iodide oxidation of<br />

B2.5 10 9 M 1 s 1 was abstracted. The data were ascribed<br />

to be most consistent <strong>with</strong> formation of ion-pairs.<br />

Since that time there have been a number of studies aimed at<br />

abstracting the rate at which the Ru II form of the sensitizer is<br />

regenerated. These experiments were usually performed by<br />

monitoring the recovery of the MLCT absorption bleach after<br />

pulsed-laser excitation at wavelengths where the iodide oxidation<br />

products did not appreciably absorb light. While this has<br />

proven to be a reasonable way of quantifying rate constants<br />

for regeneration of the Ru II state, little information regarding<br />

the mechanism(s) of iodide oxidation is obtained. For this<br />

reason, we briefly summarize the key observations.<br />

Most studies of this type were performed <strong>with</strong> N3/TiO2. At<br />

low iodide concentrations, the regeneration rate was found to<br />

be first order in iodide. At higher iodide concentrations, a<br />

static component was often observed. Under the 0.5 M iodide<br />

concentration of a DSSC, regeneration is often stated to be<br />

complete <strong>with</strong>in 10 ns. 11,13 The rate constant for regeneration<br />

of the oxidized dye, Ru III (bpy) 2(dcb)/SnO 2, by iodide was<br />

determined to be 1.2 10 10 M 1 s 1 . 356 Durrant and coworkers<br />

have recently provided evidence that the regeneration<br />

rate is dependent on the E o (Ru III/II ) of the sensitizer. 360 With<br />

Ru(dcb)2(CN)2/TiO2 thin films an intermediate was observed<br />

and assigned to a [Ru III ,I ] ion-pair. Reaction of this <strong>with</strong> a<br />

second iodide was proposed to yield I2 .<br />

The rate of reactivity of iodide <strong>with</strong> N719 + /TiO2(e ) increased<br />

in the presence of Li + and other cations <strong>with</strong> large<br />

<strong>charge</strong>-to-radius ratios. 361 It was also noted that the half-time<br />

for sensitizer regeneration abruptly shortened when the concentration<br />

of Li + was increased to between 10 and 50 mM,<br />

Fig. 31(a). Using electrophoretic measurements, the point of<br />

zero z-potential (PZZP) was determined to occur at 3 mM<br />

Li + , a concentration slightly less than that required for the<br />

abrupt change in half-time, Fig. 31(b). Also, by titration of<br />

iodide to positively <strong>charge</strong>d TiO2 particles in the presence of<br />

Mg 2+ , experimental data suggested that iodide adsorbed<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 139


Fig. 31 (A) Plot of the inverse half-lives for the regeneration of N719 + /TiO2(e ) by iodide versus the logarithm of the concentration of Li + .<br />

(B) A similar plot depicting the point of zero z-potential (PZZP) for TiO2 nanoparticles as a function of the logarithm of the Li + concentration for<br />

(b) unsensitized TiO 2 and (d) N719/TiO 2. Both plots for N719/TiO 2 noticeably change behavior near 2 10 2 –2 10 3 M. Taken from Fig. 4a<br />

and 3, respectively, of ref. 361.<br />

<strong>with</strong>in the Helmholtz layer of the particles even in the presence<br />

of bulky dyes. It was concluded that the abrupt change to<br />

faster sensitizer regeneration occurred due to ion-pairing of<br />

iodide anions <strong>with</strong> the TiO2 surface or sensitizer resulting in an<br />

increased occurrence of the faster termolecular reaction (11).<br />

ii Regeneration by donors other than iodide. An examination<br />

of the Latimer-type diagram in Scheme 1 reveals the<br />

significant problem <strong>with</strong> the I3 /I redox shuttle required for<br />

champion DSSCs. Iodide is oxidized at +1.33 V vs. NHE<br />

(or +1.03 V if reaction (11) is operative) at the sensitized<br />

electrode and I 3 is ideally reduced at +0.04 V (or I 2 at +0.21 V<br />

given the I 3 - I 2 + I equilibrium) at the Pt counter<br />

electrode. Thus at least a half of a volt of free energy is lost<br />

<strong>with</strong> this redox mediator under standard conditions. While the<br />

Latimer-type diagram depicts aqueous values, there is reason<br />

to believe that the losses are almost as large under nonstandard<br />

conditions in acetonitrile electrolytes, thus accounting<br />

for the non-optimal Vocs that are typically measured in<br />

DSSCs. Another issue <strong>with</strong> the I3 /I redox mediator is that a<br />

facile reduction of I3 at the counter electrode in DSSCs is<br />

required so as to minimize voltage losses. Platinum has a large<br />

exchange current density and <strong>transfer</strong> coefficient for this<br />

reaction but is expensive. 362 Electrode materials like graphite<br />

do not perform as well and the corrosive nature of the<br />

electrolyte towards less expensive <strong>metal</strong>s like silver or copper<br />

precludes their use. 362 Similarly, I2 has an appreciable vapor<br />

pressure at room temperature and thus extra care must be<br />

taken to ensure a thoroughly and tightly sealed solar cell. 363<br />

Therefore, there is ample reason to identify alternative redox<br />

mediators for DSSCs.<br />

a Organic donors. With a few notable exceptions there has<br />

been very little progress in using one-electron <strong>transfer</strong>, outersphere<br />

redox couples as mediators. The published literature<br />

does not accurately reflect the experimental efforts that have<br />

been put forth in this area. This stems from the fact that it is<br />

neither rewarding nor easy to publish data on solar light-toelectrical<br />

power conversion efficiencies of o0.1%. Some time<br />

ago we showed that phenothiazine donors were able to<br />

efficiently regenerate the oxidized sensitizer. 329 However, one<br />

needed a pico-ammeter to measure any photocurrent due to<br />

quantitative recombination of TiO 2(e )s <strong>with</strong> PTZ + . In other<br />

words, PTZ + molecules were unable to escape the mesoporous<br />

film before recombination. This result appears to be very<br />

general. Gregg and co-workers found similar behavior <strong>with</strong><br />

FeCp2 donors. 363 By coating the sensitized electrode <strong>with</strong><br />

silanes, a large increase in the photoelectrochemical response<br />

was observed that was reasonably attributed to attenuation of<br />

the recombination reaction of TiO 2(e )s <strong>with</strong> FeCp 2 + , Fig. 32.<br />

In early aqueous DSSC studies, Gra¨tzel showed that hydroquinone<br />

was a satisfactory donor. 364 A polycrystalline TiO2<br />

(anatase) electrode sensitized <strong>with</strong> Ru(dcb)3 2+ in 10 mM<br />

aqueous NaCl (pH 2.6 <strong>with</strong> HCl) solution <strong>with</strong> 1 mM hydroquinone<br />

gave maximum IPCE values of 44%. Three years<br />

later, a DSSC containing 1 mM aqueous HClO4 and either<br />

10 mM hydroquinone/100 mM LiClO4 or 1 M KI electrolyte<br />

resulted in similar maximum IPCE values, Fig. 33. 365<br />

A comparison of halide redox mediators in acidic aqueous<br />

electrolyte, i.e. 1 mM HClO 4, illustrated that Ru(dcb) 3/TiO 2<br />

thin-film electrodes in electrolyte solution containing 1 M<br />

LiClO 4/1 mM Br 2 resulted in a monochromatic light-toelectrical<br />

power conversion efficiency of 12%. 365 However,<br />

the redox mediator was outperformed by the I3 /I<br />

redox mediator under short-circuit conditions. The more<br />

negative photocurrent onset observed for iodide, relative to<br />

Fig. 32 Current–voltage curves under simulated solar irradiance conditions<br />

and in the dark showing that silanization of a Ru(bpy) 2(dcb)/TiO 2<br />

thin film electrode dramatically improved the current–voltage characteristic<br />

of regenerative solar cells employing FeCp 2 +/0 as the redox couple.<br />

This data is consistent <strong>with</strong> the silanes attenuating TiO2(e )+FeCp2 +<br />

recombination. Taken from Fig. 7a of ref. 363.<br />

140 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 33 Current–voltage curves for Ru(dcb)3/TiO2 in aqueous electrolyte<br />

illustrating that bromide and dihydroquinone function nearly<br />

as well as iodide in DSSCs under short-circuit conditions. Taken from<br />

Fig. 3 of ref. 365.<br />

hydroquinone and bromide, suggested a surface adsorptioninduced<br />

shift in the flatband potential in the iodide-containing<br />

electrolyte.<br />

A comparative study of the pseudohalide redox mediators<br />

(SeCN) 2/SeCN and (SCN) 2/SCN <strong>with</strong> the standard I3 /I<br />

redox mediator in N3/TiO2 regenerative DSSCs was reported.<br />

366 As the reduction potentials of the pseudohalides<br />

were 190 and 430 mV more positive than I3 /I , respectively,<br />

it was postulated that an increased Voc would result. Interestingly,<br />

the Voc for the DSSC containing the (SeCN)2/KSeCN<br />

redox mediator was about the same as for the I3 /I redox<br />

mediator while that for (SCN)2/NaSCN was considerably<br />

smaller. The iscs differed by a factor of four under monochromatic<br />

(500 nm) light excitation. The injection yields were<br />

independent of the redox mediator in 250 mM LiClO4 acetonitrile<br />

electrolyte containing 100 mM of the sodium or potassium<br />

salt of the reduced redox mediator, but the rate of<br />

regeneration followed the order I 4 SeCN 4 SCN . While<br />

SCN and SeCN oxidation by N3 + /TiO2 was thermodynamically<br />

favored, the oxidation kinetics were sluggish which<br />

allowed a larger fraction of the injected electrons to recombine<br />

<strong>with</strong> N3 + .<br />

Interestingly, Wang and Grätzel observed more promising<br />

behavior for Z907/TiO2 thin films <strong>with</strong> the (SeCN)2/SeCN<br />

pseudohalide redox mediator in the 1-ethyl-3-methylimidazolium<br />

(EMI) selenocyanate ionic liquid <strong>with</strong> added K(SeCN)3. 367<br />

Although this ionic liquid was found to be 35 times less viscous<br />

than the traditional 1-propyl-3-methylimidazolium (PMI)<br />

iodide ionic liquid, it was over 28 times more conductive at<br />

room temperature and could solubilize approximately eight<br />

times more (SeCN)2/SeCN than PMI could <strong>with</strong> I3 /I .<br />

By transient absorption spectroscopy, it was shown that<br />

Z907 + /TiO2(e ) could be regenerated fastest in EMI-SeCN<br />

as compared to PMI-I and the analogous EMI-SCN ionic<br />

liquid. This was contrary to the findings of Oskam et al. in<br />

acetonitrile electrolytes. 366<br />

It was also shown that the<br />

maximum IPCE was close to unity and the overall lightto-electrical<br />

power conversion efficiency under 1 sun,<br />

AM1.5-simulated irradiation was 7.5%.<br />

Recent studies have investigated 2,2,6,6-tetramethyl-1-piperidinyloxy<br />

radical (TEMPO) as a possible redox mediator.<br />

368,369 Nitrosyl tetrafluoroborate (NOBF 4) was added to<br />

TEMPO in order to generate a TEMPO + /TEMPO redox<br />

couple in a 1 : 9 stoichiometry, similar to the ratio of I 3 /I<br />

employed in champion DSSCs. When 1 M TEMPO was<br />

compared to the same concentration of iodide both the isc<br />

and Voc were slightly increased. The light-to-electrical power<br />

conversion efficiency for an organic sensitizer bound to TiO2<br />

under 1 sun, AM1.5-simulated irradiation was 5.4%. When<br />

the TEMPO concentration was decreased to 0.1 M, the Voc<br />

actually increased to B910 mV.<br />

b Transition-<strong>metal</strong> donors. Octahedral Co II diimine compounds<br />

have proven to be effective donors for sensitizer<br />

regeneration and Co III/II redox mediators have led to promising<br />

light-to-electrical power conversion efficiencies in DSSCs.<br />

The Co III/II self-exchange rate constants are known to be<br />

particularly sluggish, behavior that is reasonably understood<br />

by the d 6 /d 7 electronic configurations that give rise to large<br />

inner-sphere reorganization energies. 370 It is possible that<br />

these same electronic factors are responsible for the slow rate<br />

constants for TiO2(e )+Co III recombination reactions and<br />

the reasonable photocurrent efficiencies that have been reported<br />

when Co III/II redox couples have been used.<br />

The first studies of cobalt mediators were by Grätzel and coworkers.<br />

371 A DSSC based on the [Co III/II (dbbip)2] 3+/2+<br />

redox couple, where dbbip is 2,6-bis(1 0 -butylbenzimidazol-2 0 -<br />

yl)pyridine), resulted in photovoltaic performance that rivalled<br />

the traditional I3 /I redox mediator when a B150 nm thick<br />

spray-pyrolyzed titania underlayer was deposited on the electrode<br />

substrate. It was shown that the exchange current<br />

density for the Co III/II couple at fluorine-doped tin oxide<br />

(FTO) was 7 10 6 A cm 2 in an acetonitrile–ethylene<br />

carbonate (40 : 60, v/v) electrolyte. 362 As this value was at<br />

least two orders-of-magnitude lower than that measured at<br />

platinum but more than two orders of magnitude higher than<br />

that of the I3 /I redox couple measured at FTO in the same<br />

electrolyte, a titania blocking layer was employed to slow this<br />

undesirable reaction. Using the cis-Ru(4-methyl-4 0 -hexadecylbpy)(dcb)(NCS)2<br />

sensitizer and a 1 : 9 stoichiometric ratio of<br />

Co III :Co II in the same electrolyte mixture, a maximum IPCE<br />

of 465% was realized and, under 0.094 suns, AM1.5simulated<br />

irradiation, a light-to-electrical power conversion<br />

efficiency of 5.2% was measured. 371 The use of a neutral<br />

sensitizer was found to be necessary in order to attenuate<br />

the adsorption of cationic redox species onto TiO 2. When<br />

Co II (dbbip) 2 2+ was added above a threshold of 10 mM, the<br />

second-order rate constant for regeneration—first order in<br />

N3 + /TiO2 and first order in Co II (dbbip)2 2+ —was 2.9<br />

10 6 M 1 s 1 , approximately an order-of-magnitude smaller<br />

than values reported for NaI. However, at 100 mM the<br />

pseudo-first-order rate constants were similar to those found<br />

<strong>with</strong> the same concentration of TBAI. 361 The change in<br />

apparent second-order rate constant was thought to be due<br />

to surface adsorption of the cationic redox couple at low<br />

concentrations. Adding to previous work, it was shown that<br />

the i sc for a DSSC employing Z907/TiO 2 was dependent on<br />

the counterion of the solution Co III/II redox mediator; the<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 141


perchlorate salt worked the best. 372 As expected, for the<br />

perchlorate-based redox mediators whose E o (Co III/II ) varied<br />

over 190 mV, a similar 180 mV variation in V oc was realized.<br />

The largest V oc recorded was 660 mV accompanied by a 7.9%<br />

light-to-electrical power conversion efficiency under 0.1 suns,<br />

AM1.5-simulated irradiation.<br />

Upon introduction of LiClO4 to DSSCs the lifetime of the<br />

TiO2(e )s 373,374 increased for cobalt-based redox couples<br />

whereas for the I3 /I redox mediator it decreased. This was<br />

rationalized as being due to a decrease in the local concentration<br />

of the cationic or neutral cobalt-based redox couples near<br />

the TiO2 surface when cationic Li + was present. 373,374<br />

A family of cobalt redox couples employing derivatives of<br />

bpy, phen, and tpy ligands were studied by Bignozzi, Elliott,<br />

and colleagues. 375 The best cobalt-based mediator, based on<br />

[Co III/II (DTB) 3] 3+/2+ perchlorate (DTB = 4,4 0 -di-tert-butylbpy),<br />

resulted in DSSCs exhibiting light-to-electrical power<br />

conversion efficiencies <strong>with</strong>in 80% of that of a comparable<br />

I3 /I -mediated DSSC under 1 sun, AM1.5-simulated conditions.<br />

Also, in contrast to the I3 /I redox mediator, addition<br />

of Li + to the cells increased not only the isc but the Voc as well!<br />

This was proposed to be due to a decrease in the rate of<br />

recombination of TiO2(e )s and Co III most likely from an<br />

increase in overpotential for reduction of the Co III species at<br />

the back FTO contact. In addition, cyclic voltammograms<br />

<strong>with</strong> platinum electrodes revealed sluggish interfacial, Co III/II<br />

electron-<strong>transfer</strong> kinetics relative to carbon and gold.<br />

Although gold was optimal, FTO electrodes coated <strong>with</strong><br />

graphite nanoparticles initially outperformed platinum in<br />

DSSCs; however, the carbon-coated FTO electrodes degraded<br />

<strong>with</strong> time. Nevertheless, the initial response was encouraging<br />

and shows promise for replacing platinum <strong>with</strong> a less-expensive,<br />

carbon-based material for use as a counter electrode.<br />

Although large reorganization energies and slow-electron<br />

<strong>transfer</strong> kinetics for cobalt-based redox couples are advantageous<br />

as they attenuate the unwanted, recombination reaction,<br />

TiO 2(e ) + Co III - TiO 2 + Co II , these characteristics<br />

are undesirable <strong>with</strong> respect to sensitizer regeneration,<br />

S + /TiO 2(e )+Co II - S/TiO 2(e )+Co III . Rapid sensitizer<br />

regeneration and sluggish recombination kinetics are traits<br />

that make the I3 /I redox mediator optimal. By using comediators<br />

in conjunction <strong>with</strong> [Co III/II (DTB)3] 3+/2+ it was<br />

proposed that these traits could be realized in non-iodidebased<br />

systems. 124 Both PTZ and FeCp2 were employed in<br />

Z907/TiO2 DSSCs, due to their small reorganization energies,<br />

rapid electron <strong>transfer</strong> kinetics, and reduction potentials intermediate<br />

between that of [Co III/II (DTB) 3] 3+/2+ and Z907 +/0 .<br />

By transient absorption spectroscopy the bleach due to<br />

Z907 + /TiO 2(e ) was found to recover in the presence of 0.1 M<br />

donor in the order FeCp 2 4 PTZ 4 Co(DTB) 3 2+ 4 no<br />

donor, whereas by chronocoulometry at FTO, PTZ/Co II 1:2<br />

molar mixtures were found to turnover 45% faster than<br />

FeCp2/Co II mixtures. Thus, the maximum IPCE, i.e. 480%,<br />

was achieved for 0.075/0.15 M PTZ/Co II (1:2 molar ratio) in<br />

acetonitrile after generating steady currents by photolysis for<br />

10–15 min in order to generate some Co III . The Voc and FF,<br />

650 mV and 0.63, respectively, under 0.1 suns, AM1.5simulated<br />

conditions, were both larger than for an equivalent<br />

DSSC employing the LiI/I 2 redox system (0.3/0.03 M).<br />

However, the light-to-electrical power conversion efficiency<br />

was less due to mass-transport limitations of the bulky<br />

cobalt redox mediator. By binding Os(dcb) 2Cl 2 to FTO the<br />

exchange current for [Co III (DTB) 3] 3+ reduction was greatly<br />

enhanced. 376 When employed in an N3/TiO 2 DSSC, the i sc and<br />

V oc were only slightly attenuated as compared to a gold<br />

counter electrode and, using a three electrode measurement,<br />

the potential of the Os(dcb)2Cl2/FTO counter electrode was<br />

only slightly perturbed near open-circuit conditions.<br />

Co III/II redox couples based on triazine ligands have been<br />

synthesized and characterized. 377 The heteroleptic Co(triazine-R)-<br />

Cl2 compounds were shown to have E o (Co III/II )=B+0.75 V<br />

vs. SCE (+0.99 V vs. NHE 229 ), considerably more positive<br />

than previously reported Co-based redox mediators. However,<br />

this redox couple has yet to be employed in a functioning<br />

DSSC. It will be interesting to see if regeneration by this redox<br />

mediator can compete kinetically <strong>with</strong> <strong>charge</strong> recombination.<br />

Cu I has a d 10 electronic configuration and compounds like<br />

Cu(bpy)2 + often adopt a tetrahedral geometry in solution and<br />

in the solid state. The Cu II form is subject to a Jahn–Teller<br />

distortion that often manifests itself in a geometry <strong>with</strong> more<br />

co-planar diimine ligands, i.e. a flattening, and a fifth ligand<br />

from solvent or a counterion axially ligated. It is possible to<br />

photoinduce these structural changes and they have been<br />

characterized by time-resolved X-ray techniques. 378 Like the<br />

Co III/II redox mediators, Cu II/I couples have large reorganization<br />

energies, slow self-exchange rate constants and show some<br />

modest success as mediators in DSSCs. For example, Cu I -<br />

pyridyl and Cu I -pyridyl-quinoline compounds have been studied<br />

in DSSCs. 379 The best-performing mediators produced a<br />

maximum IPCE of B40% and yielded higher Vocs and FFs<br />

than the I3 /I redox couple under the same experimental<br />

conditions. This was attributed to a decreased dark current<br />

due to the large reorganization energy of the Cu II/I redox<br />

couple.<br />

Unfortunately the large reorganization energies for Cu II/I<br />

redox mediators suffer the same pitfalls as their Co III/II<br />

counterparts, i.e. slow sensitizer regeneration. Thus, a Cu I<br />

compound <strong>with</strong> a distorted tetrahedral geometry was employed<br />

in order to help reduce the large reorganization<br />

energy. 380 When [Cu II/I (dmp)2] 2+/+ was employed as the<br />

redox mediator, where dmp is 2,9-dimethyl-phen, a light-toelectrical<br />

power conversion efficiency of 2.2% under 0.2 suns,<br />

AM1.5-simulated irradiation was obtained <strong>with</strong> N719/TiO2based<br />

DSSCs. The methyl groups were proposed to prevent<br />

planarization of the dmp ligands which manifests itself in a<br />

positive shift in E o (Cu II/I ). Significantly, a higher V oc was<br />

realized <strong>with</strong> the copper mediator as compared <strong>with</strong> I 3 /I<br />

under the same experimental conditions.<br />

iii Time scale for regeneration. Rates of regeneration of the<br />

oxidized sensitizer that is produced after excited-state injection<br />

have now been quantified <strong>with</strong> organic and inorganic donors.<br />

By covalently binding the donor to the sensitizer, intramolecular<br />

regeneration is observed. Iodide oxidation can be complicated<br />

by the presence of multiple reaction mechanisms and<br />

by ion-pairing <strong>with</strong> the sensitizer or semiconductor surface,<br />

but nonetheless is now reasonably well understood.<br />

142 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 34 (A) Absorption and photoluminescence spectra of Ru(dtb)2(dcb)/TiO2 in 0.1 M LiClO4 acetonitrile electrolyte (red spectra) and in neat<br />

acetonitrile after removal of the LiClO 4 by ten neat acetonitrile washings (black spectra). (B) Transient absorption difference spectra for three<br />

Ru(dtb)2(dcb)/TiO2 thin films at the indicated surface coverages and delay times measured after pulsed 532 nm excitation in 0.1/0.5 M LiClO4/<br />

TBAI acetonitrile electrolyte. Overlaid are simulations of the data represented by dashed lines. Inset: Time-resolved, single wavelength absorption<br />

difference spectra measured at 510 nm for each surface coverage, corresponding to cation <strong>transfer</strong>, and a single difference spectrum measured at<br />

433 nm (black spectrum <strong>with</strong> orange fit), corresponding to I 3 loss due to TiO 2(e )+I 3 recombination. Taken from Fig. 1 and 2, respectively, of<br />

ref. 381.<br />

Recent results from our laboratories suggest new directions<br />

for fundamental research and raise the question of what the<br />

term ‘regeneration’ actually means. These new results are best<br />

understood <strong>with</strong> an example, [Ru(dtb)2(dcb)] 2+ , where dtb is<br />

4,4 0 -di-tert-butyl-bpy. Fig. 34(a) shows the absorption and PL<br />

spectra of a Ru(dtb)2(dcb)/TiO2 thin film immersed in 0.1 M<br />

LiClO4 acetonitrile and in neat acetonitrile. In the presence of<br />

Li + both maxima red-shifted and their intensity decreased<br />

relative to neat acetonitrile. The significant quenching of the<br />

PL results from enhanced excited-state electron injection into<br />

TiO 2 as described previously herein. 254<br />

Pulsed 532 nm excitation of Ru(dtb) 2(dcb)/TiO 2 in 0.1/0.5 M<br />

LiClO 4/TBAI acetonitrile electrolyte resulted in the microsecond<br />

absorption difference spectrum, Fig. 34(b). Under such<br />

conditions, one would expect to observe a TiO2(e ) and<br />

oxidized iodide products, e.g. I3 . Regardless of the mechanism<br />

for iodide oxidation, most of the I2 should have<br />

disproportionated by this sufficiently long time delay. The<br />

absorption features characteristic of I3 (l o 420 nm) and<br />

TiO2(e )s (l 4 560 nm) were indeed observed. However, the<br />

absorption band centered at 460 nm and the bleach at 510 nm<br />

could not be assigned to any conceivable electron-<strong>transfer</strong><br />

products.<br />

Spectral modeling indicated that the absorption features at<br />

460 and 510 nm resulted from [Ru(dtb)2(dcb)] 2+ sensitizers<br />

that were regenerated in a Li + -deficient milieu. In other<br />

words, the sensitizers that were initially photo-excited had<br />

an absorption spectrum shown in red while immediately after<br />

regeneration their spectrum was that shown in black,<br />

Fig. 34(a). Overlaid on the data in Fig. 34(b) are simulations<br />

based on the weighted addition of (1) the absorption spectrum<br />

of I3 , (2) the TiO2(e ) absorption spectrum, and (3) the<br />

difference in the absorption spectra of Ru(dtb) 2(dcb)/TiO2 in<br />

the absence minus the presence of Li + . Similar sensitizer<br />

absorption features were observed when PTZ was used in<br />

place of iodide. The excellent agreement between observed and<br />

simulated spectra provided compelling evidence that these<br />

sensitizers were regenerated in an environment that lacked<br />

outer-sphere Li + interaction(s). This behavior was also observed<br />

after pulsed-light excitation of Ru(bpy)2(dcb)/TiO2 and<br />

Ru(bpy)2(dcbq)/TiO2 in 0.5 M iodide electrolyte as well as<br />

previously in the published literature for N3/TiO2 382 and<br />

Ru(bpy) 2(dcb)/SnO2. 356 In all cases, absorption features were<br />

observed that were not due to oxidized iodide products,<br />

TiO2(e )s, or other redox states of the sensitizers. They were,<br />

however, reasonably described as sensitizers regenerated in an<br />

environment depleted of Li + .<br />

The absorption changes that correspond to cation <strong>transfer</strong><br />

were well described by the Kohlrausch–Williams–Watts<br />

(KWW) function for a distribution of rate constants (distributions<br />

are shown in Fig. 35(a)):<br />

" #<br />

b<br />

t<br />

C ¼ Co exp<br />

ð12Þ<br />

where t o is the most representative lifetime, i.e. the mode, and<br />

b is inversely related to the width of the underlying Levy<br />

distribution, 0 o b o 1. 383–385 Kohlrausch first proposed the<br />

function empirically and it was later popularized by Williams<br />

and Watts. The inverse Laplace transform is known analytically<br />

for discrete values of b and can be approximated for<br />

others allowing the distribution of rate constants to be directly<br />

recovered. 386 Values for to = 4.1 ( 2.5) 10 5 s and b = 0.16<br />

0.01 were found. The low b values corresponded to a broad<br />

Levy distribution of rate constants. Tens of microseconds to<br />

even milliseconds were required for completion of the cation<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 143<br />

to


Fig. 35 (A) Normalized Levy distributions of lifetimes calculated using the indicated values of b from the KWW model. As b approaches 1, the<br />

distribution approaches a Dirac d function in shape and thus the kinetics would begin to follow a simple first-order model. Taken from Fig. 2 of<br />

ref. 402. (B) Simulated time-resolved spectra based on the random-flight multiple-trapping model of Tachiya and colleagues on the left and the<br />

multiple-trapping, nearest-neighbor CTRW model of Nelson et al. on the right. Although similar in shape, the trapping–detrapping rate is four<br />

orders-of-magnitude slower in the former. Taken from Fig. 18 of ref. 401.<br />

<strong>transfer</strong> when the Ru(dtb)2(dcb)/TiO2 thin films contained a<br />

high surface coverage of sensitizers. While the large tert-butyl<br />

groups may inhibit cation motion, similar time scales for<br />

cation <strong>transfer</strong> were observed for Ru(bpy) 2(dcb)/TiO 2 and<br />

N3/TiO 2. The time scale for cation <strong>transfer</strong> was in itself<br />

surprising considering that the sensitized film was immersed<br />

in 0.1 M Li + -containing electrolyte, thus highlighting the<br />

locality of said effect. Even though the sensitizer surface<br />

coverage was high, only a small concentration of sensitizers,<br />

approximately equal to that of TiO2(e )s, were found to be in<br />

this Li + -deficient environment.<br />

The spectral data indicated that Li + <strong>transfer</strong> away from the<br />

sensitizer occurred in o10 ns. Furube et al. reported timeresolved<br />

infrared data consistent <strong>with</strong> picosecond Li + <strong>transfer</strong><br />

after excited-state injection by coumarin sensitizers, behavior<br />

attributed to Coulombic repulsion between the oxidized coumarin<br />

and Li + . 387 As cations are required for <strong>charge</strong> compensation<br />

of the injected electron, that could also induce Li +<br />

to migrate away from the oxidized sensitizer. As mentioned<br />

previously, intercalation of Li + is known to accompany<br />

reduction of anatase TiO2. In this regard, we found that the<br />

same sensitizer spectral changes could be observed by partial<br />

electrochemical reduction of the TiO2, i.e. when no oxidized<br />

sensitizer was present, indicating that <strong>charge</strong> compensation<br />

plays a role.<br />

After fast excited-state electron injection into TiO 2 and<br />

regeneration by iodide, sensitizers were present in an environment<br />

distinctly different from that prior to light absorption.<br />

Significantly, the newly generated sensitizers were in an environment<br />

that is known to be less favorable for excited-state<br />

electron injection. 254 Under 1 sun, AM1.5 irradiation, the slow<br />

(ms—ms) cation <strong>transfer</strong> is not expected to limit the efficiency<br />

of DSSCs as a typical Ru II sensitizer absorbs light approximately<br />

twice every second. 388 However, at higher irradiances<br />

or at planar TiO2 surfaces this effect may limit light-toelectrical<br />

power conversion efficiencies. In all cases, the sensitization<br />

rate constants shown in Fig. 1 need to be modified.<br />

The oxidized Ru III sensitizer may be reduced to Ru II on a<br />

nanosecond time scale, however it is not brought back to the<br />

environment prior to light absorption until slow (ms—ms)<br />

cation <strong>transfer</strong> has taken place.<br />

5. Charge recombination<br />

Charge-recombination processes at sensitized semiconductor<br />

interfaces have been studied in considerable detail. Reactions<br />

of TiO2(e )s <strong>with</strong>: (A) the oxidized sensitizer; and (B) acceptors<br />

in the electrolyte, Step IV in Fig. 1, have received much<br />

attention. As described in section 4/B/i/b, regeneration of the<br />

oxidized sensitizer by iodide is rapid and quantitative in<br />

champion DSSCs. Therefore, reaction IV–A is generally not<br />

relevant to these cells. It is, however, important in DSSCs<br />

employing alternative redox mediators or sensitizers whose<br />

ground-state reduction potentials are less favorable for regeneration<br />

than that of N3, E o (Ru III/II ) o +0.85 V vs. SCE 25<br />

(+1.09 V vs. NHE 229 ). It may also become relevant at high<br />

irradiances or in viscous electrolytes. The transparent nature<br />

of the mesoporous, nanocrystalline TiO2 (anatase) thin films<br />

allows this process to be quantified in a transmission mode<br />

<strong>with</strong> signal-to-noise ratios comparable to what can be<br />

achieved in fluid solution. Mechanistic insights have been<br />

gained from transient absorption measurements made under<br />

open-circuit conditions in the absence of an external electron<br />

donor such that each injected electron recombines <strong>with</strong> oxidized<br />

sensitizers.<br />

In champion DSSCs, <strong>charge</strong> recombination to acceptors<br />

<strong>with</strong>in the I3 /I electrolyte is a very inefficient process. The<br />

fraction of TiO2(e )s that recombine by this pathway is<br />

usually so small (o0.01) that it does not significantly influence<br />

the isc. However, TiO2(e )s that recombine by this pathway<br />

are thought to have a significant influence on the quasi-Fermi<br />

level of the semiconductor and hence a large effect on the Voc.<br />

Intensity-modulated photovoltage/photocurrent spectroscopy<br />

(IMVS/IMPS) and time-domain transient photovoltage/<br />

photocurrent decays <strong>with</strong> appropriate modeling, have provided<br />

some insights into the mechanisms of this unwanted<br />

reaction.<br />

A Electron-transport-limited <strong>charge</strong> recombination<br />

In the late 1990s, our group at Johns Hopkins University and<br />

the groups at Imperial College in London provided evidence<br />

that <strong>charge</strong> recombination was not slow because of inherently<br />

small rate constants but because efficient separation of the<br />

144 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


injected electron and the oxidized sensitizer resulted in nongeminate<br />

recombination. Our own kinetic data showed that<br />

<strong>charge</strong> recombination was modeled by an equal-concentration,<br />

second-order process much like the analogous process in fluid<br />

solution. 254,389 The concentration of <strong>charge</strong>-separated states<br />

was controlled by modulating the excitation irradiance or the<br />

Li + concentration. When the concentration of <strong>charge</strong>separated<br />

states, i.e. Ru III /TiO2(e ), was varied by over a<br />

factor of ten, the same second-order rate constant was<br />

abstracted from the data. A single second-order rate constant<br />

could be used to fit the first few microseconds of recombination,<br />

whereas a sum of two rate constants, i.e. a bi-second<br />

order kinetic model, was required to fit the entire transient.<br />

The possibility that a distribution of second-order rate constants<br />

had underlain the observed kinetic behavior could not<br />

be ruled out. The observed second-order rate constant, which<br />

was obtained directly from the transient absorption data, had<br />

the unconventional units of s 1 . Accurate conversion to the<br />

more common units of M 1 s 1 was complicated by the<br />

heterogeneity of the sample which resulted in an ill-defined<br />

homogeneous optical path length. Assuming adherence to<br />

Beer’s Law and an optical path length equivalent to the film<br />

thickness, the back-electron <strong>transfer</strong> rate constants for<br />

Ru III (bpy)2(dcb)/TiO2(e ) - Ru II (bpy)2(dcb)/TiO2 were<br />

found to be B9 10 11 and 3 10 10 M 1 s 1 in 1.0 M LiClO 4<br />

acetonitrile electrolyte. Other researchers have employed the<br />

same methodology to abstract a weighted-average of the<br />

equal-concentration, second-order rate constants in units of<br />

M 1 s 1 . 390 The key advance was that efficient separation of<br />

the TiO2(e ) from S + occurred thereby giving rise to somewhat-isolated<br />

TiO2(e )s and oxidized sensitizers. Recombination<br />

was second order in nature not first order as had been<br />

previously assumed in the kinetic modeling. In fact, actinometry<br />

measurements showed no evidence for first-order,<br />

geminate recombination; every injected electron recombined<br />

by the second-order mechanism.<br />

Shortly thereafter, Nelson and the group from Imperial<br />

College proposed a model for <strong>charge</strong> recombination that<br />

involved transport of the injected electron to the oxidized<br />

sensitizer. 183,185 The central idea was that <strong>charge</strong> carriers<br />

become trapped in localized states and that the kinetics for<br />

<strong>charge</strong> transport are dominated by the time constants for<br />

release from those states. Transient phenomena of this kind<br />

are termed ‘dispersive’ 391–393 when they are rate-limited by this<br />

step. A numerical model initially derived by Scher and Montroll<br />

based on a ‘continuous-time’ random walk (CTRW)<br />

where species move by diffusion on a lattice was utilized. 394,395<br />

The dispersive nature of such kinetics was introduced by<br />

applying a power-law, waiting-time distribution time step,<br />

c p t 1 b ,0o b o 1. For ‘normal’ diffusion the time step<br />

would be drawn from a Poisson distribution, c p e t/t . The<br />

result of such a model are kinetics that follow the KWW<br />

function, which represents a distribution of rate constants as<br />

shown in Fig. 35(a), 383–385 <strong>with</strong> b being equal to the inverse of<br />

the DOS non-ideality factor. 185 Also, <strong>with</strong> this model, the t1/2<br />

for the recombination ought to vary <strong>with</strong> the number of<br />

TiO2(e )s per particle, n, as<br />

t 1/2 = Cn 1/b<br />

(13)<br />

where t1/2 = to(ln 2) 1/b and C is a constant. 184,185 This has<br />

been confirmed experimentally. A mean lifetime for the kinetic<br />

process can also be calculated by the first moment of the<br />

KWW function:<br />

tKWW ¼ to<br />

b<br />

G 1<br />

b<br />

ð14Þ<br />

where G(x) is the Gamma function. 396,397 This model is often<br />

applicable in fractal systems or regarding relaxation in solids.<br />

398,399 When such systems are applied to fast <strong>charge</strong><br />

recombination from TiO2(e )s to oxidized sensitizers or acceptors<br />

in solution the model fits rather well. 178,182,184,186<br />

Two models for trap-limited diffusion in disorder media<br />

were proposed by Nelson et al. 185 The first was based on<br />

multiple-trapping-limited recombination as derived from the<br />

CTRW model where steps to all nearest neighbors were<br />

equally likely. The other was based on tunneling-limited<br />

recombination, where quantum-mechanical tunneling can<br />

result in long-range interactions, i.e. farther than nearestneighbor.<br />

Fits to experimental <strong>charge</strong>-recombination data<br />

under external bias and plots of the half-lives versus the<br />

concentration of TiO2(e )s strongly ruled out the latter model<br />

and supported the former given an exponential DOS.<br />

Later, an extension to the multiple-trapping, nearestneighbor<br />

CTRW model proposed by Nelson et al. was reported<br />

by Tachiya and colleagues. 400,401 This new randomflight<br />

multiple-trapping model included the possibility of many<br />

neighbor interactions, where the probability that the detrapped<br />

electron will be captured by any empty, surface trap<br />

state <strong>with</strong>in the nanoparticle is equal. Although similar in<br />

shape to the model proposed by Nelson et al., the calculated<br />

trapping–detrapping rate as a function of time was many<br />

orders-of-magnitude slower, Fig. 35(b).<br />

A hopping model that differs from the multiple-trapping<br />

model has also been proposed by Bisquert. 403 Instead of<br />

activated detrapping to the conduction band, electron transport<br />

occurs by direct <strong>transfer</strong> via localized states located at<br />

energies just below Ecb. However, a very high carrier density is<br />

needed to validate the model as all of the above models predict<br />

similar behavior at lower carrier densities.<br />

There is now a wide body of <strong>charge</strong>-recombination data that<br />

is well modeled by the multiple-trapping, nearest-neighbor<br />

CTRW model and the KWW function. It allows a great deal<br />

of experimental data to be modeled <strong>with</strong> only two independent<br />

variables. However, the derived parameters (to and b) do not<br />

always provide insights into the underlying dynamics. The<br />

KWW function has been ‘‘derived’’ by at least three different<br />

groups: (1) the already discussed CTRW model of Scher and<br />

Montroll; 394,395 (2) a distribution of serially linked first-order<br />

rate constants by Anderson; 404 and (3) fractal time concepts by<br />

Shlesinger. 405 Although not as rigorous, Plonka has also<br />

shown that dispersive, second-order kinetics can lead to<br />

behavior that is well modeled by the KWW function. 406<br />

Anderson’s model is based on a distribution of first-order rate<br />

constants whose magnitudes decrease <strong>with</strong> time. Such behavior<br />

can be very difficult to distinguish from a second-order<br />

process where the rate decreases <strong>with</strong> time but the rate<br />

constant does not. We have in fact shown that it is often<br />

impossible to conclude whether a distribution or a sum of two<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 145


discrete rate constants underlie complex kinetic behavior<br />

based simply on the quality of the fit. 407 The point here is<br />

that while modeling <strong>charge</strong> recombination based on Scher and<br />

Montroll’s CTRW model makes good physical sense, it is not<br />

a unique fit and other models may ultimately provide more<br />

insights.<br />

i Comparisons to single-crystal anatase. Experimentally a<br />

great deal is known about the transport of injected electrons<br />

through mesoporous, nanocrystalline TiO2 (anatase) thin<br />

films. The mobility of and diffusion coefficient for free conduction-band<br />

electrons in single-crystal anatase and rutile<br />

TiO2 in the absence of solvent were found to be on the order<br />

of 1–10 cm 2 V 1 s 1 and 10 1 cm 2 s 1 , respectively, and were<br />

inversely related to temperature due to optical phonon scattering.<br />

408–410 Similar, but often slightly lower, values for<br />

TiO 2(e )s in mesoporous, nanocrystalline TiO 2 are obtained<br />

when trapping–detrapping events are removed either from<br />

the experiment—based on the technique—or from the<br />

results—based on modeling. 411–413 Using a trap-filling model<br />

Ko¨nenkamp was able to estimate the TiO2(e ) free carrier<br />

mobility for air- or N2-filled, mesoporous, nanocrystalline<br />

TiO2 (anatase) thin film, Schottky barrier electrodes to be<br />

B2.4 cm 2 V 1 s 1 . 412 O’Regan and colleagues were able to<br />

measure TiO2(e ) mobilities for air-filled, mesoporous, nanocrystalline<br />

TiO2 (Degussa P25) thin films using terahertz<br />

spectroscopy. 411 This technique is unique as it measures the<br />

average mobility of <strong>charge</strong>s due to intraparticle transport on<br />

the B10 ps time domain, prior to electron trapping. Although<br />

the calculated mobilities were two orders-of-magnitude slower<br />

than the values calculated for single-crystal rutile electrodes,<br />

this could be explained by employing the Drude model and the<br />

appropriate Maxwell–Garnet effective medium theory. It was<br />

concluded that the reduced terahertz mobility observed in the<br />

porous sample was due to screening of the applied field by the<br />

polar TiO2 matrix. Additionally, Bisquert and colleagues<br />

determined that the trapping–detrapping-limited TiO2(e )<br />

diffusion reported on an average, effective diffusion coefficient<br />

whereas, for comparison to single-crystal values, the more<br />

appropriate tracer, or jump, diffusion coefficient should be<br />

obtained. 180,413–416 By employing the analytical solutions to<br />

this novel model, Peter determined tracer diffusion coefficients<br />

near Voc conditions that were comparable to those obtained<br />

for single-crystal anatase, i.e. about one order-of-magnitude<br />

smaller. 413<br />

ii Ambipolar diffusion. An interesting aspect of diffusion<br />

that is rather unique to nanocrystalline TiO2 (anatase) thin<br />

films in DSSCs results from the high ionic concentrations and<br />

mesoporosity of the thin film electrodes. Diffusion of <strong>charge</strong>d<br />

particles in solution and in highly conductive media are<br />

shielded by counterions as required in order to maintain<br />

‘quasi-neutrality.’ Thus, over large volumes neutrality is preserved,<br />

however on the scale of the Debye length <strong>charge</strong><br />

imbalances can exist. For nanocrystalline, anatase TiO2 the<br />

situation is often different as the pores of anatase TiO2 are<br />

large enough to accommodate cations <strong>with</strong> large <strong>charge</strong>-toradius<br />

ratios 156,190,205,207–210,212–215,417–420 and thus the Debye<br />

length is on the order of 1 A˚ . 410 In champion DSSCs each<br />

injected electron is thought to be immediately shielded by a sea<br />

of oppositely <strong>charge</strong>d Li + and the long-range, macroscopic<br />

electric field across the film is negligible, Fig. 36. 18,159–162<br />

It is for this reason that TiO 2(e ) diffusion is governed by<br />

the concerted motion of the electronic and electrolyte <strong>charge</strong>s<br />

per the formula: 17,410,421<br />

Damb ¼ DnDpðn þ pÞ<br />

ð15Þ<br />

nDn þ pDp<br />

where Damb is this ambipolar diffusion coefficient and n, Dn, p<br />

and Dp are the anionic- and cationic-<strong>charge</strong> densities and<br />

diffusion coefficients, respectively. 219 As champion DSSCs<br />

employ 0.5 M electrolyte solutions and are often evaluated<br />

under 1 sun, AM1.5 conditions where the concentration of<br />

TiO2(e )s is far less than the concentration of electrolyte in<br />

solution, Damb is approximately the diffusion coefficient for<br />

the less dense carrier, i.e. the TiO2(e )s, even though the<br />

apparent diffusion coefficient for TiO2(e )s, Dn, is larger. This<br />

large concentration of oppositely <strong>charge</strong>d and mobile Li +<br />

results in the injected electron being the minority carrier. 410<br />

The net outcome is an attenuation in the rate of TiO2(e )<br />

diffusion and an increase in the rate of Li + diffusion relative to<br />

their rates of diffusion in each other’s absence. 17<br />

This ambipolar diffusion model was described in the early<br />

1950s and solved computationally using a non-linear model in<br />

the late 1960s. 422,423 Searson and co-workers first reported<br />

that the TiO2(e ) diffusion coefficient, in thin-film electrodes,<br />

was dependent on the light intensity and thus also on the<br />

concentration of TiO2(e )s. 424 Hagfeldt and colleagues later<br />

reported that TiO2(e ) diffusion in unsensitized mesoporous,<br />

nanocrystalline TiO2 (anatase) thin-film electrodes followed a<br />

cation-dependent mechanism. 425 They termed the solution<br />

Fig. 36 (A) A diagram illustrating the space-<strong>charge</strong> potential drop across a TiO 2 nanocrystallite that is B20 nm in diameter before and after<br />

contact <strong>with</strong> a solution electrolyte. (B) Color-coded topographical model illustrating the relative potential distribution for an ordered mesoporous<br />

network of such nanocrystallites. Taken from Fig. 5 and 6, respectively, of ref. 18.<br />

146 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


electrolyte an image cloud care of classical physics terminology.<br />

426 Reduction of the electrolyte concentration from 500 to<br />

20 mM resulted in a five-fold decline in the diffusion coefficient,<br />

indicative of an ambipolar diffusion mechanism. With<br />

laser pulses at 2 Hz in 500 mM LiClO 4 acetonitrile electrolyte,<br />

the diffusion coefficients were shown to significantly slow,<br />

possibly due to insufficient time for Li + deintercalation from<br />

<strong>with</strong>in the anatase lattice. These same researchers later showed<br />

similar behavior for sensitized N3/TiO2 thin-film electrodes. 427<br />

Although not fit to an ambipolar diffusion model it was<br />

apparent that the cation concentration limited the TiO2(e )<br />

collection time.<br />

Frank and colleagues were the first to quantify ambipolar<br />

diffusion coefficients for mesoporous, nanocrystalline TiO 2 (anatase)<br />

thin-film electrodes. 410 Using laser-pulsed photocurrent<br />

transients, both in the absence and presence of a constant<br />

background illumination, ambipolar diffusion coefficients for<br />

N719/TiO2 electrodes were obtained over a large range of<br />

excitation energies. These diffusion coefficients ranged greatly<br />

from 3 10 8 to 10 4 cm 2 s 1 for low to high irradiances,<br />

respectively. The difference in the calculated ambipolar diffusion<br />

coefficient and TiO2(e ) diffusion coefficient was greatest under<br />

the highest illumination intensities studied; however the difference<br />

in the values was only B15%.Thus,TiO2(e ) diffusion<br />

coefficients were satisfactory estimates for the more accurate<br />

ambipolar diffusion coefficients. Further support for the ambipolar<br />

diffusion model was later established by these same<br />

researchers based on the arrival-time detection of current from<br />

TiO2(e )s at the back FTO contact, rather than displacement<br />

current, under counter-electrode side illumination. 428 This implied<br />

that the electric field of the TiO2(e )s was shielded by the<br />

electrolyte and that only TiO2(e )s that physically arrived at the<br />

FTO–TiO2 junction registered a photocurrent. It was originally<br />

thought that the large concentration of electrolyte in functioning<br />

DSSCs would effectively shield the TiO2(e )s and have little<br />

effect on their transport. However, this is only the case under<br />

steady-state conditions. At early times after a perturbation, the<br />

diffusion coefficient for TiO 2(e )s is substantial and an ‘ionic<br />

drag’ on the free electron mobility is present. 16 It was shown that<br />

when the sea of counter-<strong>charge</strong>d species was rather dilute the<br />

TiO2(e ) transport became less dispersive at early times. In<br />

contrast, the concentration of counter-<strong>charge</strong>d species had little<br />

influence on the steady-state limit of the TiO2(e )s diffusion<br />

coefficient. Thus, on short time scales the ambipolar effect<br />

hindered fast electron transport through the TiO2 film while<br />

under steady-state conditions, where transport was trap limited,<br />

ionic drag was generally absent.<br />

Yanagida and co-workers found evidence that further supported<br />

the ambipolar diffusion model using Li + -concentrationdependent<br />

transient photocurrent studies <strong>with</strong> unsensitized<br />

TiO2 thin-film electrodes in ethanol electrolyte. It was shown<br />

that the ambipolar diffusion model adequately described the<br />

data and that the TiO2(e ) diffusion coefficient was over<br />

two orders-of-magnitude larger than that of Li + . 429 Additional<br />

studies performed by varying the irradiance in either<br />

700 mM or 5 mM LiClO4-electrolyte solutions resulted<br />

in the expected ambipolar diffusion trend. In 700 mM electrolyte,<br />

a direct relationship between the ambipolar diffusion<br />

coefficient and irradiance was observed while in 5 mM electrolyte,<br />

an inverse relationship existed as expected by the diffusion<br />

of Li + now being rate limiting, Fig. 37(a). Using similar<br />

experimental procedures, these same researchers showed that<br />

the diffusion coefficients for various cations, i.e. Li + ,Na + ,<br />

Mg 2+ , TBA + , dimethylhexylimidazolium cation (DMHI + ),<br />

could accurately be extracted from the low-concentrationelectrolyte<br />

data fit to the ambipolar diffusion model. 430<br />

However, at higher electrolyte concentrations significant and<br />

unexpected increases in the ambipolar diffusion constant were<br />

obtained. Only for the TBA + data did the ambipolar diffusion<br />

model result in a satisfactory fit for all concentrations studied,<br />

Fig. 37(b). The empirically determined diffusion coefficients<br />

for the other cations were well above expected values in the<br />

order DMHI + 4 Li + 4 Na + , assumed to be due to specific<br />

adsorption. By employing UV-Vis spectroscopy of the soaking<br />

solutions, it was indeed shown that there was multilayer<br />

absorption of DMHI + on TiO2 whereas practically no<br />

absorption of TBA + occurred. Additionally, when plots<br />

the ambipolar diffusion coefficient versus the concentration<br />

of cation were obtained for Li + and TBA + , a noticeable<br />

hysteresis was present for Li + assigned to Li + adsorption<br />

onto TiO 2.<br />

iii Activation energy. As stated above, the diffusion coefficient<br />

for TiO2(e )s is dependent on their concentration. 424<br />

Thus, it would seem likely that the activation energy for<br />

electron transport <strong>with</strong>in this network of nanocrystallites<br />

Fig. 37 (A) A log–log plot of the empirical diffusion coefficients as a function of TiO2(e ) density. The monotonic increasing trend for the data in<br />

700 mM LiClO 4 and the inflection in the 5 mM data can be satisfactorily modeled by the ambipolar diffusion model. Taken from Fig. 5 of ref. 429.<br />

(B) A nearly perfect fit of the ambipolar diffusion constant versus the logarithm of the concentration of TBA + data to the ambipolar diffusion<br />

model for a large concentration of TiO 2(e )s (circles); also shown is data at low TiO 2(e ) density (squares). Taken from Fig. 3 of ref. 430.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 147


would also be TiO2(e )-concentration dependent. By employing<br />

conductivity measurements, activation energies for electron<br />

transport of B0.3 eV were obtained for mesoporous,<br />

nanocrystalline TiO 2 thin-film electrodes under typical DSSC<br />

working conditions. 431<br />

Most measurements of transport times in mesoporous,<br />

nanocrystalline TiO2 thin-film electrodes are determined under<br />

short-circuit conditions, as elsewhere transport is often RC<br />

limited. 432 However it is often desirable/necessary to determine<br />

these transport times near DSSC working conditions, i.e.<br />

near VPP/Voc. 413 O’Regan and colleagues have developed a<br />

novel method for doing so by measuring photovoltage transients<br />

at various preset voltages and entering the results in a<br />

zero-free-parameter model. 432 Using this novel method it was<br />

shown that TiO 2(e ) transport though N3/TiO 2 thin-film<br />

electrodes fits the multiple-trapping model, whereby detrapping<br />

limits TiO 2(e ) transport and recombination. 433 The<br />

method also allowed for a more accurate calculation of<br />

activation energies for TiO2(e ) transport, as prior methods<br />

did not allow for the facile correction of the temperature<br />

dependence of Ecb. The results seemed to indicate that the<br />

activation energy was not as large as the energy difference<br />

between the trap states and the conduction band, consistent<br />

<strong>with</strong> other reports. 434–436 Thus the results are in agreement<br />

<strong>with</strong> the aforementioned hopping model proposed by Bisquert<br />

where electrons need not fully thermalize to the conduction<br />

band in order to hop between trap states <strong>with</strong>in TiO 2. 403<br />

Following the work of Bisquert and Vikhrenko, 415,416 a<br />

model employing a quasi-static approximation was proposed<br />

that accounted for previously measured activation energies<br />

<strong>with</strong>out invoking the temperature dependence of Ecb. 437 In<br />

contrast, this model predicted concerted and equal shifts in Ecb<br />

and the energy of the quasi-Fermi level <strong>with</strong> temperature.<br />

B Recombination to the oxidized sensitizer<br />

Early studies of TiO 2 colloidal solutions and thin films sensitized<br />

to visible light <strong>with</strong> Ru II -based and organic sensitizers<br />

established that <strong>charge</strong> recombination to the oxidized sensitizer<br />

occurred on a tens- to hundreds-of-microseconds time<br />

scale. 158,438 The observation of efficient sensitization from<br />

compounds <strong>with</strong> very short excited-state lifetimes, such as<br />

[cis-Ru(dcb)2(H2O)2] 2+ , indicated that excited-state electron<br />

injection was a sub-nanosecond process. 439 The question<br />

then naturally arose: why does such a fortuitous difference<br />

in interfacial <strong>charge</strong>-separation and <strong>charge</strong>-recombination<br />

rate constants exist at the TiO 2 interface? For MLCT excited<br />

states part of the explanation was that injection occurred<br />

from the p* orbitals of a surface-bound, dcb ligand while<br />

recombination was to the t2g orbitals of the Ru III -<strong>metal</strong><br />

center. In other words, there is a built-in type of rectification<br />

in these sensitizers whose orbitals provide strong electronic<br />

coupling for <strong>charge</strong> separation but inhibit recombination.<br />

218,219 It was also known that <strong>charge</strong> recombination<br />

was in the Marcus inverted region whereas excited-state injection<br />

was nearly activationless. 438 Such orbital participation<br />

and thermodynamics could explain the large difference in<br />

interfacial <strong>charge</strong>-separation and <strong>charge</strong>-recombination rate<br />

constants.<br />

i Lateral electron <strong>transfer</strong> via surface-bound adsorbates.<br />

While it is often tacitly assumed that transport of the injected<br />

electron is most relevant to <strong>charge</strong> recombination, it is important<br />

to emphasize that the oxidized sensitizer also has some<br />

mobility. Lateral <strong>charge</strong> <strong>transfer</strong> across semiconductor surfaces<br />

is often initiated by <strong>charge</strong>-<strong>transfer</strong> reactions at the<br />

transparent conductive electrode (TCE) that supports the<br />

TiO2 thin film. Such hole <strong>transfer</strong> can almost entirely be<br />

eliminated <strong>with</strong> the addition of a blocking layer on the back<br />

TCE support prior to thin-film deposition. Thus the major<br />

means for hole <strong>transfer</strong> is by lateral sensitizer-mediated hopping<br />

of <strong>charge</strong> or physical movement/diffusion of the bound<br />

sensitizers. Should the diffusion coefficient for such a process<br />

be independent of sensitizer concentration, the latter mechanism<br />

is assumed. However, a sharp surface-coverage onset to<br />

the diffusion is consistent <strong>with</strong> an underlying self-exchange,<br />

<strong>charge</strong>-<strong>transfer</strong> reaction and a percolation threshold. 71 A<br />

percolation threshold is formally defined for the conductivity<br />

inside a composite material as ‘‘the critical concentration<br />

above which an infinite cluster of conductive sites spans the<br />

network’’. 71 It is often assessed by measuring diffusion coefficients<br />

over a range of surface coverages via chronoamperometry/chronocoulometry<br />

and Cottrell/Anson plots (i vs.t 0.5 /Q<br />

vs. t 0.5 ). However a novel technique utilizing chronoabsorption<br />

measurements and spectrophotometric Anson plots<br />

(DA vs. t 0.5 ) has also been utilized. 71<br />

Electrochemical investigation of the E o (Ru III/II ) for sensitizers<br />

bound to mesoporous, nanocrystalline TiO 2 (anatase)<br />

thin-film electrodes revealed that by integration of the area<br />

under the cyclic voltammogram, B10% of the concentration<br />

of spectroscopically quantified sensitizers had been oxidized/<br />

reduced. 72 Although some dyes could directly adsorb to the<br />

FTO electrode, this could not wholly explain the B10% that<br />

were electro-active as this corresponded to over an order-ofmagnitude<br />

larger surface coverage than was physically possible.<br />

It was proposed, for the first time, that self-exchange<br />

<strong>charge</strong> <strong>transfer</strong> processes across the TiO 2 surface could be<br />

occurring.<br />

In the first study of lateral hole <strong>transfer</strong> across the surface of<br />

mesoporous, nanocrystalline <strong>metal</strong>-oxide thin films, a percolation<br />

threshold was found to exist. 71 This threshold was found<br />

to be 50% of saturation surface coverage for phosphonated<br />

triarylamines adsorbed onto TiO2, ZrO2, orAl2O3 as determined<br />

by chronoabsorption measurements and spectrophotometric<br />

Anson plots. The mechanism for this hole <strong>transfer</strong> was<br />

deduced to be via self-exchange hole <strong>transfer</strong> <strong>with</strong> eventual<br />

mediation by the back TCE support. This process was not<br />

limited by the ion motion in solution for the systems studied.<br />

Employing Z907 bound to <strong>metal</strong>-oxide, thin-film electrodes,<br />

it was found that a chemically reversible anodic wave was<br />

present on TiO2 and highly insulating Al2O3. 440 The percolation<br />

threshold was found to be B50% of saturation surface<br />

coverage and faster lateral hole diffusion coefficients were<br />

observed in acetonitrile-based electrolytes versus purely ionic<br />

liquids. This was proposed to be due to the two orders-ofmagnitude<br />

higher viscosity of the ionic liquid that resulted in<br />

slower effective ambipolar-diffusion rates. It was also clearly<br />

deduced that the increased hole diffusion coefficients for<br />

Z907 + and [cis-Ru(dmb)(dcb)(NCS) 2] + under saturation<br />

148 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


surface coverages were due to efficient hole <strong>transfer</strong> between<br />

isothiocyanate ligands. This was evident by comparison to the<br />

slower diffusional, hole-hopping rates for (NC ) 2-containing<br />

dyes, Ru(bpy) 3 2+ -type dyes, dyes <strong>with</strong> trans isothiocyanate<br />

ligands, cis-Ru(dmb)(dcb)(NCS) 2 <strong>with</strong> mercury-poisoned isothiocyanate<br />

ligands, and N3—whose intermolecular-isothiocyanate<br />

ligands are further separated on the surface. The<br />

fastest diffusion coefficient for hole <strong>transfer</strong> was achieved<br />

<strong>with</strong> [cis-Ru(dmb)(dcb)(NCS)2] + and was determined to be<br />

1.1 10 8 cm 2 s 1 .<br />

While studying the percolation threshold for three TiO2bound<br />

sensitizers containing donor moieties shown in Fig. 24,<br />

it was discovered that one of them exhibited a long-lived<br />

photochroism. 330 Photochroism occurs when oxidized-donor<br />

spectral features are still present even when oxidation of the<br />

sensitizer is followed by application of an appropriate bias that<br />

could thermodynamically reduce the oxidized donors but not<br />

the TiO2 DOS. This was attributed to the extended conjugation<br />

of the sensitizer that inhibited its free rotation and the<br />

formation of a surface-conducting monolayer. Thus, facile<br />

percolation of <strong>charge</strong> from the back FTO contact to every<br />

surface-bound donor could not be realized.<br />

The percolation threshold for Os(bpy)2(dcb)/TiO2 thin-film<br />

electrodes was found to be 60% of saturation surface coverage.<br />

73 However, at coverages less than the percolation threshold,<br />

but <strong>with</strong> the addition of Ru(bpy) 3 2+ in solution,<br />

mediation of oxidative hole <strong>transfer</strong> occurred upon stepping<br />

the potential positive of E o (Ru III/II ). Similarly, addition of<br />

Os(bpy)3 2+ in solution increased the rate of both reductive<br />

electron <strong>transfer</strong> and oxidative hole <strong>transfer</strong> upon biasing the<br />

film in the appropriate direction due to solution-based, dyemediated<br />

<strong>charge</strong> <strong>transfer</strong>.<br />

By employing a sensitizer <strong>with</strong> a lower E o (Ru II/+ ) than Ecb<br />

and relying on ‘hot’-electron injection, laser-flash photolysis of<br />

Ru II (bpy)2(dcbq)/TiO2 thin films in acetonitrile resulted in<br />

immediate, i.e. o10 ns, formation of Ru III (bpy) 2(dcbq)/TiO 2<br />

and Ru II (bpy) 2(dcbq )/TiO 2 <strong>charge</strong>-separated states. 135 Fits<br />

to stretched exponentials via the KWW function resulted in<br />

average rate constants, i.e. t o 1 , for [Ru II (bpy)2(dcbq )] + +<br />

[Ru III (bpy)2(dcbq)] 3+ recombination of 8 5 10 5 s 1 . But a<br />

question remained: was recombination predominantly due to<br />

electron or hole <strong>transfer</strong> reactions? By employing chronoabsorption<br />

measurements and spectrophotometric Anson plots,<br />

the diffusion coefficient for the Ru II/+ self-exchange reaction<br />

on TiO2 was found to be over an order-of-magnitude larger<br />

than that for the Ru III/II reaction. Solution self-exchange rate<br />

constants for Ru(bpy) 3 2+ were <strong>with</strong>in a factor of two the same<br />

for Ru III/II and Ru II/+ in acetonitrile. 441–444 It was proposed<br />

that the TiO 2 DOS mediated electron, but not hole, <strong>transfer</strong>.<br />

Using Ru II (bpy) 2(dcbq)/TiO 2, Fe III (PPIX)Cl/TiO 2 and<br />

Fe III (PPIX)(py)2/TiO2 thin-film electrodes, in acetonitrile,<br />

methanol, and methanol, respectively, no percolation threshold<br />

for electron <strong>transfer</strong> was observed in TBA + electrolytes<br />

(PPIX is protoporphyrin IX and py is pyridine). 445 Interestingly,<br />

after being reduced the diffusion coefficients for their<br />

oxidation were over two orders-of-magnitude slower for the<br />

Fe II -based, PPIX coordination compounds. The diffusion<br />

coefficient for electron-<strong>transfer</strong> reduction for each molecular<br />

acceptor was at an intermediate value between these two<br />

re-oxidation extremes but were <strong>with</strong>in experimental error of<br />

one another. This was rationalized based on a Gerischer-type<br />

model where the fluctuating energy levels for Fe II had a much<br />

poorer overlap <strong>with</strong> the TiO 2 DOS as compared to those of<br />

dcbq .<br />

ii Final TiO2(e )–sensitizer + <strong>charge</strong> recombination<br />

a E o (M +/0 ) and TiO2 DOS and quasi-Fermi-level dependence.<br />

Semiclassical, non-adiabatic Marcus Theory predicts a<br />

parabolic-dependence on the logarithm of the electron-<strong>transfer</strong><br />

rate constant <strong>with</strong> the standard-state driving force for the<br />

reaction. 240,446,447 The maximum rate constant occurs at the<br />

vertex of this parabola and represents activationless electron<br />

<strong>transfer</strong> and thus should be temperature independent.<br />

Electron-<strong>transfer</strong> processes occurring at larger driving forces<br />

are actually slower and are located in an energetic/kinetic<br />

region termed the inverted region. Moser and Gra¨tzel reported<br />

practically temperature-independent rate constants for <strong>charge</strong><br />

recombination from colloidal TiO2 to surface-bound, oxidized,<br />

organic sensitizers over a 4200 degree temperature<br />

window. 438 Based on numerical simulations employing a<br />

quantum-mechanical model for non-adiabatic electron <strong>transfer</strong>,<br />

including an average high-frequency vibrational mode<br />

from the sensitizer, 305–310 it was shown that under conditions<br />

of moderate solvent reorganization energy, practically activationless<br />

electron-<strong>transfer</strong> behavior could be observed well into<br />

the Marcus inverted region. Thus, this highly exergonic recombination<br />

reaction was concluded to fall deeply <strong>with</strong>in the<br />

Marcus kinetic inverted region even though the roughly<br />

temperature-independent rate constants eluded to activationless<br />

behavior.<br />

Driving-force-dependent electron <strong>transfer</strong> can be quantified<br />

at dye-sensitized TiO2 interfaces where excited-state electron<br />

injection into TiO2 leaves an electron at a particular standardstate<br />

potential and a ‘‘hole’’ on the sensitizer. The E cb, and<br />

subsequently the free energy of the TiO 2(e )s, can be varied by<br />

altering the pH or the concentration of cations, as was employed<br />

in excited-state injection studies in section 3/B/ii, while<br />

the free energy of the ‘‘hole,’’ i.e. E o (Ru III/II ), can be controlled<br />

through synthetic manipulation of the coordinated ligands.<br />

Lever has an empirical model that allows these potentials to<br />

be accurately determined before the sensitizer is synthesized. 448<br />

However, in some cases the environment and the proximity of<br />

the sensitizer to the TiO2 surface can in itself result in different<br />

measured E o (Ru III/II )s. Zaban et al. have previously shown that<br />

the E o (Ru III/II )oftheRu II (LL)(mpt)CN sensitizer, where LL is<br />

1,2-bis(4 0 -methyl-bpy-4-yl)ethane and mpt is 4 0 -phosphonic<br />

acid-tpy, became pH-dependent when bound to mesoporous,<br />

nanocrystalline TiO 2 (anatase) thin films and shifted in a nearly<br />

Nernstian fashion in concert <strong>with</strong> Ecb. 313 Similar behavior was<br />

also shown for eight other sensitizers who possessed pHindependent<br />

E o (M +/0 )s in fluid solution but whose E o (M +/0 )<br />

shifted 21 to 53 mV/pH unit when bound to TiO2. 312 These<br />

sensitizers were either organic or inorganic <strong>with</strong> <strong>metal</strong> center<br />

being Ru II ,Fe III or Mg II and ligands being tetracarboxyphthalocyaninato-,<br />

dcb- or dpb-based, where dpb is 4,4 0 -diphosphonic<br />

acid-bpy. It was proposed that the position of the adsorbed<br />

sensitizer <strong>with</strong>in the ionic double layer could explain the<br />

differences in shifts of E o (M +/0 ) per pH unit.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 149


Fig. 38 Plot of the logarithm of the inverse of the half-times/lifetimes versus the driving force for two different studies of TiO 2(e )+S +<br />

recombination. The trend in (A) implies that the electron <strong>transfer</strong> was in the Marcus inverted kinetic region while the trend in (B) seems to follow<br />

activationless electron <strong>transfer</strong>. Taken from Fig. 5 of ref. 390 and Fig. 4 of ref. 450, respectively.<br />

Employing a family of MLCT sensitizers based on osmium,<br />

ruthenium and rhenium whose ground state reduction potential<br />

varied by about B960 mV, it was shown that recombination<br />

kinetics from a TiO2(e ) to an oxidized sensitizer were independent<br />

of sensitizer employed. 389 The transient data was<br />

successfully fit to an equal-concentration, bi-second-order<br />

kinetic model on a 100 ns and longer time scale. The data<br />

could also be successfully fit to four first-order rate constants<br />

but the rationale for this fit was less apparent. Of note was that<br />

the raw transient data was insensitive to the sensitizer reduction<br />

potential, the molecular geometry, the nature of the <strong>metal</strong><br />

center employed—i.e. Re, Ru, or Os—and the number of<br />

carboxylic acid groups present—i.e. two or four. The insensitivity<br />

of the second-order rate constants to these parameters<br />

was attributed to the reaction being rate limited by TiO2(e )oxidized<br />

sensitizer (h + ) encounters or lack of change in the<br />

apparent driving force for the recombination reaction due to<br />

concerted E cb and ground-state E o shifts, as was shown previously.<br />

312,313 Charge transport <strong>with</strong>in the sensitized film may<br />

also have been a second-order process.<br />

In a separate study by Lewis and co-workers, <strong>charge</strong><br />

recombination to five Ru III or Os III compounds were fit to<br />

the same equal-concentration, bi-second-order recombination<br />

model. 390 By using the weighted average of the second-order<br />

rate constants in conjunction <strong>with</strong> semiclassical Marcus theory,<br />

<strong>charge</strong> recombination to oxidized sensitizers, like N3 + /TiO2,<br />

was found to fall in the Marcus inverted kinetic region <strong>with</strong> a<br />

total reorganization energy, l = B1.0 eV, Fig. 38(a). The<br />

temperature-dependent electron-<strong>transfer</strong> kinetics were similar<br />

to those observed by Dang and Hupp <strong>with</strong> Ru-phen based<br />

coordination compounds electrostatically bound to colloidal<br />

SnO 2 nanoparticles. 449 The kinetics suggested that while<br />

nuclear tunneling was negligible, solvent reorganization and<br />

low-frequency, <strong>metal</strong>–ligand vibrational modes assisted the<br />

recombination reaction, as opposed to high-frequency,<br />

igand-based vibrational modes. Lewis and co-workers also<br />

reported that the activation energy for <strong>charge</strong> recombination<br />

was slightly larger for Os II - versus Ru II -based sensitizers.<br />

More recently, eight different sensitizers whose E o (Ru III/II )<br />

spanned B500 mV were employed to examine the drivingforce<br />

dependence on <strong>charge</strong> recombination. 450 The transient<br />

spectroscopic data was fit to a multiple-trapping, nearestneighbor<br />

CTRW kinetic model by using the KWW function<br />

and the driving-force dependence was quantified based on t1/2,<br />

the time it took for half of the injected electrons to recombine.<br />

Using the inverse of these half-lives the data was shown to be<br />

relatively insensitive to variations in E o (Ru III/II ). This was<br />

interpreted as being indicative of reactions lying near the peak<br />

of the Marcus free energy curve, DG o = Bl, and <strong>with</strong> l =<br />

B0.8 eV, Fig. 38(b). Therefore, these authors concluded that<br />

<strong>charge</strong> recombination to N3 + , and other similar sensitizers,<br />

was nearly activationless. It is interesting to note that while the<br />

E o (Ru III/II ) values were generally in good agreement <strong>with</strong><br />

Lewis and colleagues, 390 the magnitude of the driving force<br />

differed significantly due to discrepancies in the reducing<br />

power of the TiO2(e )s, Fig. 38.<br />

Employing [Ru II (depb)3] 2+ or [Ru II (dpb)3] 10 , where depb<br />

is 4,4 0 -diethylphosphonate-bpy bound to mesoporous, nanocrystalline<br />

TiO 2 (anatase) thin films, the pH dependence of<br />

recombination in aqueous solution was studied by Hupp and<br />

co-workers. 451 The fast exponential component to the biphasic<br />

recovery was shown to be invariant of pH (or H 0) over a<br />

19 pH-unit range even though the Ecb of TiO2 is known to shift<br />

in a nearly Nernstian fashion <strong>with</strong> pH, Fig. 39(a). One might<br />

expect the E o (Ru III/II ) to shift in a concerted fashion as<br />

observed by Zaban et al., 312,313 however this was not the case<br />

as the E o (Ru III/II ) of the surface-bound sensitizer was shown to<br />

have only a minor pH-dependence (5 mV/pH unit) over a<br />

47.5 pH unit range, Fig. 39(b). 452 This less than Nernstianorder-of-magnitude<br />

shift in E o (Ru III/II ) should have been<br />

largely overcome by the potential shift in E cb. As a continuation<br />

of this study, Ru(dpb) 2(LL)/TiO 2 thin films were shown<br />

to exhibit minor, but apparent, Marcus normal region behavior,<br />

where LL were bpy and phen derivatives. 453 This unexpected<br />

result, given the large variations in driving force, was<br />

explained as sequential electron- and proton-<strong>transfer</strong> reactions.<br />

It was proposed that the rate-limiting step was backelectron<br />

<strong>transfer</strong>, however this step did not release all of the<br />

free energy in the overall reaction, and thus the variation in<br />

driving force for this step was solely dependent on changes in<br />

E o (Ru III/II ), Fig. 39(c). As the E o (Ru III/II ) differed only slightly<br />

<strong>with</strong> pH, and was actually found to vary <strong>with</strong> the z-potential<br />

150 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 39 (A) Plot of the logarithm of recombination rate constant versus pH for TiO 2 thin films. As the energy of the conduction band edge, E cb,<br />

shifts in a nearly Nernstian fashion <strong>with</strong> pH (dashed line), this plot illustrates driving-force independent recombination rates (solid line <strong>with</strong><br />

points). Taken from Fig. 4 of ref. 451. (B) Plot of the standard-state reduction potential of the surface-bound compounds (E o (Ru III/II )) versus the<br />

pH, confirming that the E 0 (Ru III/II ) did not change <strong>with</strong> the pH of the solution and thus further supporting the driving-force independent<br />

mechanism. Taken from Fig. 5 of ref. 452. (C) Pourbaix-type diagram of the TiO 2(e )+S + recombination reaction. The driving-force<br />

independent recombination rates were rationalized as being due to rate-limiting electron <strong>transfer</strong> followed by exergonic proton <strong>transfer</strong>. Taken<br />

from Fig. 5 of ref. 453.<br />

of the TiO2, only a small deviation in driving force was<br />

actually realized.<br />

The described <strong>charge</strong>-recombination studies <strong>with</strong> altered<br />

thermodynamic driving force provide a somewhat conflicting<br />

series of conclusions: rate constants independent of the<br />

ground-state reduction potential of the sensitizer, in the<br />

inverted region, and near the top of the Marcus curve.<br />

Undoubtedly, these differences arise from details of the preparation<br />

of the TiO2 thin films, surface chemistry such as TiCl4<br />

pretreatments, 454 and species present in the electrolyte. Additional<br />

research is required to identify the critical variables,<br />

especially in non-aqueous electrolytes.<br />

While the driving-force dependence for <strong>charge</strong> recombination<br />

to the oxidized sensitizer in mesoporous, nanocrystalline<br />

thin films is not completely established, the behavior in<br />

aqueous colloidal solutions appears to be more clear. Hupp<br />

and co-workers monitored <strong>charge</strong> recombination to six different<br />

Fe(CN) 5(L) n<br />

sensitizers bound to colloidal TiO2 in<br />

aqueous pH 2 solution on a nanosecond time scale. 304 The<br />

recombination kinetics were found to be biphasic <strong>with</strong> a<br />

fast first-order component followed by longer non-exponential<br />

kinetics. The first-order component varied <strong>with</strong> E o (Fe III/II )<br />

as expected if the process were occurring in the Marcus<br />

inverted region. Likewise, recombination reactions from<br />

electrostatically bound Ru II - and Os II -polypyridyl coordination<br />

compounds to colloidal SnO2 nanoparticles exhibited<br />

a pH-sensitive kinetic behavior. 455 This is in stark contrast<br />

to the pH-independent behavior observed for recombination<br />

from similar compounds covalently anchored to mesoporous,<br />

nanocrystalline TiO2 (anatase) thin films. It was<br />

suggested that this difference was due to the significant trap<br />

density in TiO2 and that ionic attachment introduced fewer<br />

surface traps.<br />

As <strong>with</strong> studies on <strong>charge</strong> separation of section 3/B/ii, it is<br />

also useful to study <strong>charge</strong> recombination under steady-state,<br />

working conditions at short circuit. The first account of such a<br />

study was performed by Gra¨tzel and co-workers in 1990. 158 By<br />

exciting at least a quarter of the sensitizers on Ru(dcb) 3/TiO2 thin-film electrodes in LiClO4 aqueous electrolytes it was<br />

shown that the half-times for the non-exponential recombination<br />

kinetics decreased by almost three orders of magnitude<br />

under forward-bias versus reverse-bias conditions. This same<br />

increase in rate was found for N3/TiO 2 thin-film electrodes in<br />

ethylene carbonate–propylene carbonate (1 : 1) solution as the<br />

irradiance was increased to generate B1 toB50 TiO 2(e )s/<br />

particle. 456 The kinetics were fit to the multiple-trapping,<br />

nearest-neighbor CTRW model and KWW function. The<br />

half-times were invariant on irradiance when o1 TiO2(e )/<br />

particle was generated.<br />

Durrant and co-workers showed the half-time for<br />

N3 + +TiO2(e ) recombination exhibited an exponential dependence<br />

on applied voltage in 0.1 M TBA + trifluoromethanesulfonate<br />

ethanol electrolyte, Fig. 40(a). 457 When immersed<br />

in ethanol electrolytes and on applying an electrochemical bias<br />

from +100 to 600 V vs. Ag/AgCl, the half-time for the<br />

recombination reaction decreased by more than seven orders<br />

of magnitude while the injection yield remained unchanged. 456<br />

By changing the electrolyte from ethanol/TBAT (Electrolyte<br />

A), where TBAT is tetrabutylammonium triflate, to CH3CN/<br />

TBAClO4/LiClO4 (Electrolyte B) and then adding 4-tertbutylpyridine<br />

(Electrolyte C), similar bias-dependent,<br />

TiO2(e ) recombination rates were observable but, again, only<br />

under conditions of 41 TiO2(e )/particle, Fig. 40(b). However,<br />

when comparing Electrolyte A to B at different biases,<br />

their half-times for recombination differed to varying degrees,<br />

but <strong>with</strong> the half-time in Electrolyte B always being larger. As<br />

the driving force for recombination would decrease in the<br />

presence of Li + due to the positive shift in Ecb, this data does<br />

not fit that of Marcus inverted behavior. Additionally, the<br />

multiple-trapping, nearest-neighbor CTRW model for recombination<br />

was further supported over a bi-second-order model.<br />

254,389,390 The greater than seven orders-of-magnitude<br />

decrease in half-time would result in approximately the same<br />

increase in initial rate. Given the proposed second-order<br />

process, the differential rate law v = k2[TiO2(e )][S + ] would<br />

require the generation of 10 7 TiO 2(e )s/particle which was<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 151


Fig. 40 (A) Bias-dependent, time-resolved, single-wavelength absorption difference spectra for N3/TiO2 thin film electrodes indicating that as the<br />

density of acceptor states in the TiO2 electrode were filled, the rate of TiO2(e )+N3 + recombination increased. Applied biases are indicated on the<br />

figure. Taken from Fig. 2 of ref. 457. (B) Plot of the logarithm of the recombination half-life versus the applied bias in the indicated electrolytes. See<br />

text for electrolyte compositions. Taken from Fig. 6 of ref. 456.<br />

highly improbable based on the magnitude of spectroscopic<br />

signals and approximate extinction coefficients.<br />

Although the rate of <strong>charge</strong> recombination of TiO2(e )s and<br />

N3 + was dependent on the concentration of electrochemically<br />

generated TiO2(e )s, this was not the case for regeneration by<br />

the solution redox mediator. 382 At moderate LiI concentration,<br />

i.e. 30 mM, and at an applied bias approximately equal to the<br />

Voc under 1 sun, AM1.5-simulated conditions, the kinetic<br />

partitioning between TiO 2(e )s + N3 + recombination and<br />

regeneration of N3 + by the reduced solution redox mediator<br />

were similar. But under the typical DSSC conditions of 0.5 M<br />

LiI and 0.05 M I2 and 1 sun, AM1.5-simulated irradiance, the<br />

rate of sensitizer regeneration would be greatly favored and<br />

TiO2(e )s + N3 + recombination would not limit performance.<br />

b Distance dependence. Distance-dependent electron tunneling<br />

behavior can be studied for recombination from a<br />

TiO2(e ) to an oxidized surface-bound acceptor. Rate<br />

constants due to tunneling should exhibit an exponential<br />

dependence on distance, eqn (8), as was the case for<br />

electron-injection tunneling behavior in section 3/B/iii.<br />

Three novel sensitizers bound to TiO2 via an mpt ligand and<br />

containing a nearby triarylamine donor moiety were studied in<br />

order to determine the distance dependence for recombination,<br />

Fig. 24. 330 The donor nitrogen atoms were calculated to<br />

be 12 A˚ ,18A˚ , and 24 A˚ from the surface. Assuming the<br />

typical dampening factor, b = 1.2 A ˚ 1 , for ‘through space’<br />

electronic coupling between the TiO2(e )s and NAr3 + , the<br />

expected 6 A ˚ -distance dependence was not observed and was<br />

off by a factor of three. It was thought that this was due to the<br />

fact that the dyes may not bind exactly in a perpendicular<br />

orientation to the surface.<br />

Employing three Ru II sensitizers containing a bpy ligand<br />

derivatized <strong>with</strong> zero, one or two oligo(xylylene) linkers,<br />

the distance dependence of TiO 2(e ) recombination was<br />

studied. 267 By comparing the weighted average of the<br />

bi-second-order rate constants 390 differences in backelectron-<strong>transfer</strong><br />

rate constants were found to be <strong>with</strong>in<br />

experimental error of one another. As was the case <strong>with</strong><br />

electron injection, the lack of expected large differences in rate<br />

constant were proposed to result from variable Ru–TiO2<br />

distances due to the flexibility of the linker groups.<br />

The distance dependence of back-electron <strong>transfer</strong> was<br />

examined in acetonitrile for three rigid-rod Ru(bpy)3 2+ -based<br />

sensitizers containing zero, one, two, or three rigid phenyleneethynylene<br />

linkers covalently attached to TiO 2 via two methyl<br />

ester anchoring groups. 268,458 It was shown that recombination<br />

kinetics from a TiO2(e )toRu III were successfully fit by<br />

an equal-concentration, bi-second-order kinetic model<br />

254,389,390 and that the average of the rate constants was<br />

essentially independent of sensitizer employed. Thus, the expected<br />

distance dependence for the rate of back-electron<br />

<strong>transfer</strong> was not observed. This was partially supported by<br />

the fact that sensitizers lacking binding groups were still found<br />

to bind to TiO2 indicating that expected Ru–TiO2 distances<br />

may be incorrect due to alternative binding orientations.<br />

Durrant and colleagues studied Ru(4,4 0 -(R) 2-bpy)(dcb) 2/TiO 2<br />

thin films, where R contained oligo(NPh 3) groups at<br />

varying distances from the Ru II -<strong>metal</strong> center, to determine<br />

the distance dependence for back-electron <strong>transfer</strong>. 332 After<br />

combining this data <strong>with</strong> other data from their laboratories,<br />

253,331,450,459 it was clearly evident that the TiO2(e )<br />

back-electron <strong>transfer</strong> rate constants displayed an exponential<br />

dependence on spatial separation <strong>with</strong> dampening factor,<br />

b = 0.95 0.2 A˚ 1 . Additionally, cis-Ru(4,4 0 -vinyl(NPh3)nbpy)(dcb)(NCS)2<br />

(n = 1, 2) were synthesized and DFT<br />

calculations suggested that their HOMOs resided predominantly<br />

on the NPh 3 moieties at a distance of 10.5 and 11.6 A˚ ,<br />

respectively. 460 When compared to Z907/TiO 2, these dyes<br />

exhibited larger t 1/2, i.e. 0.22, 1.8, 3 ms for Z907, n =1,<br />

and n = 2, respectively, that followed the expected<br />

exponential trend.<br />

152 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 41 (A) Time-resolved, single-wavelength absorption difference spectra for Ru(4 0 -PO3 2 -tpy)(NCS)3/TiO2 thin films illustrating that the rate<br />

of recombination was inversely related to the size of the Al 2O 3 overlayer illustrated in Fig. 16(a). Al 2O 3 overlayer thickness in nanometers are<br />

shown. (B) Plot of the natural logarithm of the recombination half-life versus the Al2O3 coating thickness illustrating tunneling behavior. Taken<br />

from Fig. 8 and 9, respectively, of ref. 271.<br />

Using the mono-phosphonated version of the ‘black dye,’<br />

[Ru(mpt)(NCS)3] 3 , bound to the same TiO2/Al2O3 core–shell<br />

particles mentioned above for electron injection, Fig. 16(a),<br />

the distance dependence of TiO2(e ) recombination was<br />

studied. 271 Tunneling was required for electron recombination<br />

and the kinetics were non-exponential <strong>with</strong> half-times ranging<br />

from 6 ms to 60 ms for 0 to 6 nm thick Al 2O 3 overlayers,<br />

Fig. 41(a). Using signal half-times it was shown that <strong>charge</strong><br />

recombination of TiO 2(e )s/Al 2O 3 and [Ru(mpt)(NCS) 3] +<br />

resulted in a dampening factor, b = 0.15 A˚ 1 , Fig. 41(b).<br />

This result, combined <strong>with</strong> the predominant dampening factor<br />

for injection, b = 0.11 A ˚ 1 , illustrates that recombination is<br />

attenuated to a greater degree than injection when TiO2/Al2O3<br />

core-shell designs are employed. The approximately six-timessmaller<br />

dampening factor when compared to the results from<br />

the molecular approach used by Durrant and co-workers<br />

suggests that intra-bandgap states <strong>with</strong>in the Al2O3 coating<br />

could be present and mediating this back-electron <strong>transfer</strong>.<br />

As mentioned above, Haque, Durrant, and colleagues significantly<br />

increased the <strong>charge</strong>-separated lifetime to over 4 s by<br />

employing a Ru(4,4 0 -(R) 2-bpy)(dcb) 2/TiO 2 system, where R<br />

contained a poly(vinyl-NPh3) group of about 100 units,<br />

Fig. 25(b). 332 The half-times for the <strong>charge</strong>-separated state<br />

increased from 350 ms to 5 ms to 4 s as the number of vinyl-<br />

NPh3 subunits increased from 1 to 2 to 100, respectively. Also,<br />

when fit to the multiple-trapping, nearest-neighbor CTRW<br />

model and the KWW function, the dampening factor ranged<br />

from 0.4 to 0.9 to 1, respectively; the latter indicating a monoexponential,<br />

first-order recombination mechanism, Fig. 42.<br />

Clifford et al. analyzed two free-base porphyrin sensitizers,<br />

meso-5,10,15,20-tetrakis(4-carboxyphenyl)porphyrin and<br />

meso-5-(4-carboxyphenyl )-10,15,20-tris(4-diphenylaminophenyl)porphyrin,<br />

by transient absorption spectroscopy and<br />

found that TiO2(e ) recombination <strong>with</strong> the oxidized porphyrin<br />

ring was sufficiently slowed for the latter sensitizer. 461 The<br />

former sensitizer’s kinetics were satisfactorily fit to the multiple-trapping,<br />

nearest-neighbor CTRW model and the KWW<br />

function, b = 0.31, whereas the kinetics for the latter sensitizer<br />

were perfectly first order in nature, i.e. b = 1, Fig. 43(a). This<br />

was believed to be due to a different rate-limiting step for<br />

Fig. 42 Time-resolved, single-wavelength absorption difference spectra<br />

for Ru(4,4 0 -(R)2-bpy)(dcb)2/TiO2 thin films, where R contained<br />

one triphenylamine (NPh 3) group (1), two NPh 3 groups (2), or a<br />

poly(vinyl-NPh3) group of about 100 units (3), as depicted in<br />

Fig. 25(b). Taken from Fig. 2 of ref. 332.<br />

recombination between each oxidized sensitizer and the<br />

TiO 2(e )s. While detrapping was rate limiting for the former<br />

dye, the final TiO 2(e )+S + electron-<strong>transfer</strong> step limited the<br />

latter. The explanation for a first-order rate constant, and not<br />

an equal-concentration, second-order rate constant, was that<br />

the intrinsic concentration of TiO2(e )s was much larger than<br />

the additional concentration resulting from sensitizer, excitedstate<br />

injection. 182 Thus, as the concentration of oxidized<br />

sensitizers decreased due to TiO2(e )+S + recombination,<br />

the TiO2(e ) concentration changed little resulting in the<br />

observed pseudo-first-order kinetic behavior.<br />

The back-electron <strong>transfer</strong> kinetics for two Ru II -based<br />

sensitizers were compared: N719 and a novel sensitizer <strong>with</strong><br />

a covalently bound triarylamine donor where the hole was<br />

efficiently <strong>transfer</strong>red away from TiO2 <strong>with</strong>in the instrument<br />

response time. 331 As was seen by Clifford et al. the kinetics of<br />

the faster back-electron <strong>transfer</strong> from N719 were dispersive<br />

while the slower kinetics for the novel sensitizer were first<br />

order in nature.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 153


Fig. 43 Time-resolved, single-wavelength absorption difference spectra for two free-base porphyrin sensitizers, meso-5,10,15,20-tetrakis-<br />

(4-carboxyphenyl)porphyrin (i, dye 1) and meso-5-(4-carboxyphenyl)-10,15,20-tris(4-diphenylaminophenyl)porphyrin (ii, dye 2), bound to TiO 2<br />

thin films illustrating that the recombination kinetics were dispersive, and could be fit to the multiple-trapping, nearest-neighbor CTRW model of<br />

Nelson et al., and a first-order kinetic mechanism, respectively. (A) The data is displayed as the logarithm of the DAbsorbance versus the time to<br />

illustrate the first-order nature of the latter sensitizer’s kinetics, i.e. ii, dye 2. Taken from Fig. 1(b) of ref. 461. (B) The same data displayed as the<br />

DAbsorbance versus the logarithm of the time to illustrate the dispersive nature of the former sensitizer’s kinetics, i.e. i, dye 1. Taken from Fig. 1(a)<br />

of ref. 461. (C) The same data and plot as in B but fit equally as well to the new Coulomb-trap, random-flight multiple-trapping model of Tachiya<br />

and colleagues. Taken from Fig. 4 of ref. 462.<br />

Tachiya and colleagues proposed the Coulomb trap model<br />

for back-electron <strong>transfer</strong> to oxidized sensitizers which employs<br />

the random-flight multiple-trapping model for TiO2(e )<br />

transport. 462 The model is based on the idea that a TiO2(e )<br />

only feels the Coulombic attraction to a dye cation when it is<br />

trapped near the cation. This stabilization interaction effectively<br />

increases the activation energy for the detrapping not<br />

only of said TiO2(e ) but of neighboring TiO2(e )s as well. By<br />

employing an exponential TiO2 DOS that are pre-filled to a<br />

reasonable dark concentration of TiO2(e )s, i.e. 0.1 per particle,<br />

166 and assuming that the effective dielectric constant at<br />

the site of this Coulombic attraction is low, i.e. o5, the<br />

experimental data of Clifford et al. can be modeled extremely<br />

well. Fig. 43(b)/(c) illustrates the comparison of the two<br />

models: the multiple-trapping, nearest-neighbor CTRW model<br />

by Nelson et al. in the middle, b, and the new Coulomb-trap,<br />

random-flight multiple-trapping model by Tachiya and colleagues<br />

on the right, c.<br />

C Recombination to acceptors in solution<br />

i Proposed mechanisms for recombination in I3 /I -<br />

containing electrolytes. Given the rather high iodide concentration<br />

and fast regeneration kinetics observed in champion<br />

DSSCs, the recombination reaction of TiO 2(e )s <strong>with</strong> surfacebound<br />

oxidized sensitizers, TiO 2(e )+S + , is considerably<br />

slower than sensitizer regeneration, S + +D- S+D + . 382<br />

This implies that the primary TiO 2(e ) acceptors in such cells<br />

are oxidized forms of the redox mediator in solution, D + .<br />

Molecular identification of the preferred solution acceptor<br />

species is of great importance. Possible candidates include I ,<br />

I2 ,I3 or I2, but unambiguous identification has not been<br />

obtained. Durrant, Nelson, and co-workers have reported<br />

evidence that the reaction of unsensitized TiO2(e )s <strong>with</strong> I2<br />

is kinetically faster than the reaction <strong>with</strong> I3 . 463<br />

Measurements of V oc provide an indirect, but powerful, tool<br />

for characterizing <strong>charge</strong> recombination processes. Although<br />

the V oc is the maximum Gibbs free energy that one can obtain<br />

from a regenerative solar cell, it can often be kinetically<br />

derived under steady-state, illumination conditions. The modified<br />

diode equation is often found to be relevant for DSSCs:<br />

Voc ¼ mkBT<br />

e<br />

¼ mkBT<br />

e<br />

ln Rate<strong>charge</strong><br />

in<br />

Rate <strong>charge</strong><br />

out<br />

0<br />

!<br />

Ioaf inj<br />

B<br />

ln@P<br />

ni Q<br />

ki½TiO2ðe ÞŠ<br />

i<br />

j<br />

½AijŠ nij<br />

1<br />

C<br />

A ð16Þ<br />

where m is the DOS non-ideality factor, kB is Boltzmann’s<br />

constant, T is the temperature, e is the elementary <strong>charge</strong>, Io is<br />

the incoming light flux, a is the absorptance, f inj is the<br />

injection yield, k i are the rate constants for recombination of<br />

TiO 2(e )s <strong>with</strong> acceptor species, A ij, of order v ij. 464–468 At<br />

room temperature, this equation predicts a 59 mV increase<br />

in Voc per 10-fold increase in the ratio of the rate of electron<br />

injection into TiO2 and the rate of recombination. Each<br />

decade of change in Io may also result in a 59 mV change in<br />

Voc. Decadic deviations are attributed to m a 1. Changes in<br />

the Voc measured under steady-state time and frequency<br />

(IMVS) domains are often observed by tuning the p* levels<br />

of the sensitizer, 469,470 chemisorption (for example, by 4-tertbutylpyridine)<br />

or surface functionalization, I 2-sensitizer coordination,<br />

altering the concentration or nature of cations in the<br />

electrolyte, or laterally <strong>transfer</strong>ring the hole further from the<br />

TiO 2 surface. These have been linked, directly, to changes in<br />

the recombination rate constant.<br />

An early application of this to DSSCs was observed <strong>with</strong> the<br />

intramolecular <strong>charge</strong> separation as previously described, 328<br />

where the Voc was measured for Ru(dmb)(dcb)2/TiO2<br />

and Ru(4-CH3, 4 0 -CH2-PTZ-bpy)(dcb)2/TiO2 thin film electrodes<br />

as a function of irradiance in 0.1 M LiClO4 or NaI/I2<br />

(0.5/0.05 M) propylene carbonate electrolytes. The potentials<br />

measured versus aAg + /Ag pseudo-reference electrode indicated<br />

that the V oc was 175 10 mV larger for Ru(4-CH 3,<br />

4 0 -CH 2-PTZ-bpy)(dcb) 2/TiO 2 versus Ru(dmb)(dcb) 2/TiO 2.<br />

154 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Fig. 44 Plot of Voc versus the logarithm of the irradiance for three<br />

TiO 2-bound, Ru(bpy) 3 2+ -based sensitizers containing one (circles) or<br />

two (triangles) rigid phenyleneethynylene linkers covalently attached<br />

to a 1,3,5,7-tetraphenyladamantane core or simply a deeb ligand<br />

(squares) in 0.1/0.005 M TBAI/I2 dichloromethane electrolyte. The<br />

farther the Ru II -<strong>metal</strong> center, and thus proposed ion-paired I and/or<br />

I3 , from the TiO2 surface, the longer the lifetime of the <strong>charge</strong>separated<br />

state and thus increased V oc. Taken from Fig. 3 of ref. 471.<br />

The photocurrents were approximately the same for both<br />

sensitizers over four orders-of-magnitude irradiance indicating<br />

that the numerator of eqn (16) was sensitizer independent<br />

while the denominator was not. The <strong>charge</strong> recombination rate<br />

constants were measured spectroscopically, k =3.9 10 6 s 1<br />

and k = 3.6 10 3 s 1 , respectively. With these rate constants<br />

a DVoc of B200 mV was calculated which is close to the value<br />

175 10 mV that was measured experimentally. Remarkably,<br />

these molecular interfaces behaved as ideal diodes, m =1,<br />

over this four orders-of-magnitude change in irradiance. Interestingly<br />

in the presence of the redox mediator, i.e. I 3 /I ,<br />

the non-ideality factor was 2. The V oc remained larger for<br />

Ru(4-CH 3, 4 0 -CH 2-PTZ-bpy)(dcb) 2/TiO 2 suggesting that<br />

<strong>charge</strong> recombination to oxidized iodide products was also<br />

inhibited for this compound. More recently, we have obtained<br />

experimental evidence that there may indeed be an advantage<br />

in oxidizing iodide farther from the TiO2 surface. In dichloromethane<br />

electrolytes containing 0.1/0.005 M TBAI/I2, the<br />

Voc was found to be directly related to the distance the hole<br />

was from the surface. 471 Using [Ru(bpy)2(deeb)] 2+ or tripodal,<br />

Ru(bpy) 3 2+ -based sensitizers containing one or two rigid<br />

phenyleneethynylene linkers covalently attached to a 1,3,5,7tetraphenyladamantane<br />

core and attached to the TiO 2 surface<br />

in a tripodal-shaped binding configuration, the distance between<br />

the Ru III -<strong>metal</strong> center and the TiO2 surface was<br />

varied. 268,269,471 As ion-pairing was previously shown to occur<br />

in dichloromethane <strong>with</strong> [Ru(bpy)2(deeb)] 2+ and I or<br />

I3 , 133,357,472 it was proposed that after ion-paired I photooxidation,<br />

TiO2(e ) recombination may occur to ion-paired<br />

I3 at a fixed distance from the surface. Assuming the injection<br />

yield was invariant on the length of the spacer, the Voc data<br />

highly supported this distance-dependent recombination mechanism,<br />

Fig. 44.<br />

In early studies <strong>with</strong> a series of cis-Ru(dcb)2X2/TiO2<br />

thin film electrodes (M = Ru, Os), the intensity-dependent<br />

photocurrent and V oc values could generally be rationalized<br />

based on the redox properties of the compound. The V oc<br />

was never found to be sensitizer dependent. However,<br />

recently there is some evidence that the recombination rate<br />

constants can be tuned through the p* levels of the sensitizer<br />

and halide coordination sites on macrocyclic compounds.<br />

285,470 Indeed any time large photocurrents and small<br />

Voc values are measured, it is of interest and suggests that there<br />

remains some <strong>charge</strong>-recombination pathway to the sensitized<br />

interface. Arakawa and co-workers, and more recently<br />

Bignozzi and colleagues, have identified such conditions <strong>with</strong><br />

Ru II and Os II compounds that have diimine ligands <strong>with</strong><br />

low-lying p* levels. 469,470 Recall, too, that if E cb is raised,<br />

the photogenerated TiO 2(e )s can be fully trapped on<br />

dcbq ligands, 299 behavior that is consistent <strong>with</strong> that<br />

reported here.<br />

However, in order to calculate the Voc, the absorbed photon<br />

flux, Ioa, the DOS non-ideality factor, m, the injection yield,<br />

finj, and the TiO2(e ) recombination rate based on the overall<br />

recombination mechanism—the denominator <strong>with</strong>in the<br />

Napierian logarithm—need be determined. 464–468 For the<br />

reasons previously discussed, there are no accepted and<br />

straightforward means of measuring the denominator as even<br />

the chemical identity of the acceptor is unknown, much less<br />

the mechanism by which it reacts. Doing this is no simple task<br />

given the numerous possible reaction intermediates, as depicted<br />

in Scheme 1 in section 4/B/i/a. The elementary reaction<br />

steps that have been proposed thus far are:<br />

I +I2 " I3 (9, restated)<br />

I 2 + TiO 2(e ) - I 2<br />

(17)ww<br />

2I2 - I3 +I (18a)zz<br />

I2 + TiO2(e ) - 2I (18b)yy<br />

I 2 " 2I ads<br />

(19)<br />

Iads + TiO2(e ) - I (20)<br />

TiO 2(e ) trapped " TiO 2(e ) free<br />

TiO2(e )free - TiO2(e )reactive<br />

(21)<br />

(22)<br />

X + TiO2(e )reactive - X (23)<br />

Many possible differential rate laws can be obtained by<br />

assuming that various steps are in pre-equilibrium, others are<br />

kinetically rapid, and one is rate determining. The integrated<br />

ww An equivalent reaction would be I2 + TiO2(e ) - Iads +I .<br />

zz Based on the products of (17)ww, an equivalent reaction would be<br />

2(I ads +I ) - I 2 +2I - I 3 +I .<br />

yy Based on the products of (17)ww, an equivalent reaction would be<br />

(I ads +I ) + TiO 2(e ) - 2I .<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 155


ate laws based on a range of hypothesized mechanisms are as<br />

follows (see Table 1):<br />

Rate / ½I3 Š2½TiO2ðe ÞŠ 2<br />

½I Š 2<br />

ð24aÞ<br />

Rate / ½I 3 Š½TiO2ðe ÞŠ 2<br />

½I Š<br />

Rate / ½I 3 Š½TiO2ðe ÞŠ<br />

½I Š<br />

Rate / ½I 3 Š0:5 ½TiO2ðe ÞŠ<br />

½I Š 0:5<br />

ð24bÞ<br />

ð24cÞ<br />

ð24dÞ<br />

Rate /½TiO2ðe ÞŠ ð24eÞ<br />

As is apparent, unambiguous determination of the integrated<br />

rate law and overall reaction mechanism requires<br />

knowledge of the reaction order for not just one, but two<br />

species. Thus, certain experiments alone cannot irrefutably<br />

establish the overall mechanism. Unfortunately, the current<br />

body of literature paints a somewhat conflicting story and thus<br />

results from the literature will be explained according to which<br />

of the five integrated rate laws and reaction mechanisms above<br />

could apply, i.e. eqn (24a–e).<br />

a Electrochemical techniques. Peter and colleagues studied<br />

N3/TiO2 DSSCs in acetonitrile electrolyte by IMVS/IMPS, a<br />

novel potentiostatic–galvanostatic–potentiostatic (PGP)<br />

method, and transient photovoltage/photocurrent measurements<br />

while under background illumination. Based on an<br />

inverse-square root relationship of the TiO2(e ) lifetime, tn,<br />

0.51<br />

and background light intensity, i.e. tn p Io , it was<br />

deduced that the recombination reaction of TiO2(e )s <strong>with</strong><br />

the I3 /I redox mediator was second order in TiO2(e )<br />

density, i.e. 1/0.51, and thus an I2 intermediate was proposed.<br />

223,473 This is indicative of mechanism (24a) or (24b).<br />

The same conclusion was drawn from the PGP method where<br />

the DSSCs were illuminated <strong>with</strong> a blue-light-emitting diode<br />

under open-circuit conditions and then were rapidly short<br />

circuited for chronocoulometric measurements. 474–476 The<br />

pseudo-second-order rate constant from the IMVS and PGP<br />

measurements was determined to be 0.6 and 1.1 10 4 M 1 s 1<br />

at 50 mM I3 , respectively. Unfortunately, a non-inversesquare<br />

root relationship between tn and the isc at various<br />

0.62<br />

green-light laser irradiances, i.e. tn p isc , was also determined<br />

by transient photovoltage/photocurrent measurements.<br />

477 Explanations of the results were that either<br />

recombination is not first order in TiO2(e )s or the interfacial<br />

electron-<strong>transfer</strong> rate constant depends on trap occupancy<br />

and/or the rate of TiO2(e ) diffusion. An explanation for the<br />

apparent discrepancy among experimental results may be the<br />

variations in the background/initial photon fluxes employed:<br />

effectively o0.2, 1 and o0.01 suns, AM1.5-simulated conditions,<br />

respectively. In none of these studies was the order of the<br />

reaction <strong>with</strong> respect to I2 , I3 , I or I2 concentration<br />

explored.<br />

Frank and co-workers studied N3/TiO2 DSSCs in acetonitrile–<br />

NMO (50 : 50 wt%) electrolyte, where NMO is 3-methyl-<br />

2-oxazolidinone, and by two procedures found a second-order<br />

dependence on I 3 : plots of V oc versus the concentration of I 3<br />

and IMVS in the presence of two different I 3 /I concentrations.<br />

464,478,479 The light-intensity, power-law dependence of<br />

the lifetime, as determined by IMVS and IMPS, was modeled<br />

to be the order of the reaction in TiO2(e )s and was approximated<br />

to be second order based on the calculated powers of<br />

B2.2 and B2.7, respectively. The results are consistent <strong>with</strong><br />

mechanism (24a). These same authors studied N719/TiO2<br />

DSSCs in 3-methoxypropionitrile electrolyte and deduced that<br />

recombination was TiO2(e )-diffusion limited. 480 These conclusions<br />

were based on a model of TiO 2(e )-diffusion-limited<br />

recombination and fits to transient photovoltage and photocurrent<br />

data as a function of background white-light intensity<br />

ranging approximately three orders-of-magnitude in irradiance<br />

up to B1 sun. This data was more consistent <strong>with</strong><br />

mechanism (24c), (24d) or (24e).<br />

By measuring the photovoltage under conditions of up to<br />

0.82 suns illumination and in the presence of varied concentrations<br />

of I3 for N3/TiO2 DSSCs in nitrile electrolyte,<br />

Hagfeldt and co-workers determined the recombination reaction<br />

to be first order in I3 . 481 They proposed the alternative<br />

mechanism (24b) <strong>with</strong> elementary steps (17)ww and (18b).yy<br />

Although the reaction order <strong>with</strong> respect to the density of<br />

TiO 2(e )s was not determined, mechanism (24c) was unlikely<br />

due to the high irradiances employed. The difference between<br />

these results and those of Frank and co-workers was rationalized<br />

by the solvent composition: 0 versus 50 wt% NMO,<br />

respectively, where it was proposed that NMO prevented<br />

reaction (17)ww from occurring, in favor of reaction (17).<br />

The diode-equation non-ideality factor was calculated to be<br />

very close to one, i.e. 1.08, but could be in error based on<br />

somewhat condition-dependent values often cited in the literature,<br />

i.e. B1.7–3.8. 171,178,183–185,223,224 Such error could be<br />

due to the fact that the non-ideality factor was obtained from<br />

V oc versus light intensity plots and was calculated based on the<br />

unlikely assumptions that the injection yield and TiO 2(e )<br />

recombination rate were independent of the irradiance. 254 Due<br />

to the inverse relationship between non-ideality factor and<br />

reaction order in I3 for this model, a non-ideality factor of 2<br />

would have resulted in a reaction order in I3 of 0.5, consistent<br />

<strong>with</strong> mechanism (24d).<br />

b Spectroscopic techniques. Hagfeldt and co-workers employed<br />

nanosecond transient absorption spectroscopy to study<br />

iodide turnover and regeneration from ‘black dye’/TiO 2 thin<br />

films in 3-methoxypropionitrile containing 0.5 M LiI and<br />

50 mM I2. 482 The reduction of ‘black dye’ + /TiO2 by iodide<br />

resulted in the formation of I2 , which occurred in o20 ns.<br />

At higher irradiances, the transient spectroscopic signal for<br />

I2 and TiO2(e )s at 760 nm decayed <strong>with</strong> a half-time of<br />

B100 attributed to TiO2(e ) + I2 recombination<br />

(eqn (18b)). This suggested mechanism (24b) was highly likely.<br />

The first phase of the overall biphasic kinetics were fit to an<br />

equal-concentration, second-order integrated rate law <strong>with</strong> a<br />

non-zero baseline, but <strong>with</strong> much error in the extracted rate<br />

constant. The equal-concentration, second-order kinetics were<br />

156 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


Table 1 Reaction orders for the proposed rate laws<br />

Order in:<br />

Eqn [I3 ] TiO2(e ) [I ] Comments<br />

(24a) 2 2 2 Reaction (9) is in pre-equilibrium, reaction (17) is fast, and reaction (18a) is the rate-determining step.<br />

(24b) 1 2 1 Reaction (9) is in pre-equilibrium, reaction (17) is fast, and reaction (18b) is the rate-determining step.<br />

(24c) 1 1 1 Reaction (9) is in pre-equilibrium and reaction (17) is the rate-determining step.<br />

(24d) 0.5 1 0.5 Reaction (9) is in pre-equilibrium, reaction (19) is fast, and reaction (20) is the rate-determining step.<br />

(24e) 0 1 0 Reaction (21) is rate determining and reactions (22) and (23) follow.<br />

believed to be present due to equal concentrations of I2 and<br />

TiO2(e )s formed after pulsed-light excitation via reaction (11)<br />

or equal concentrations of I and TiO2(e )s formed via reaction<br />

(10) followed by rapid and quantitative I2 formation via<br />

reaction (25):<br />

I +I - I2 (25)<br />

Based on the absorption at 760 nm at the end of the first phase<br />

of the kinetics and the extinction coefficients of TiO2(e )s and<br />

I2 , it was calculated that B1 TiO2(e ), and thus 1 I2 , was<br />

present per particle regardless of the initial equal concentrations<br />

of the species. Under these conditions or at low irra-<br />

was found to be rather slow, i.e. on<br />

diances, the decay of I 2<br />

the microsecond time scale, and likely occurred via the dismutation<br />

reaction (18a). The fact that the difference in these<br />

mechanisms occurred <strong>with</strong> irradiance and TiO2(e ) concentration<br />

did not seem coincidental as other researchers had seen<br />

similar behavior at 41 TiO2(e )/particle. 483 This implied that<br />

at the local concentration of I2 produced by the higher laser<br />

irradiances, I2 is the favorable TiO2(e ) acceptor, whereas at<br />

lower irradiances, I2 or I3 are the favorable electron acceptors.<br />

It was proposed that the loss of TiO2(e )s at this low<br />

concentration should follow a single-exponential kinetic model<br />

as now the rate-limiting step would be reaction (17), and<br />

thus mechanism (24c) would be operative. Although,<br />

one could envision that under these conditions TiO 2(e )<br />

detrapping could limit transport whereby mechanism (24e)<br />

would then explain the kinetics.<br />

Nelson et al. predicted a sublinear power-law variation of<br />

TiO2(e ) density <strong>with</strong> light intensity, i.e. n = CIo b , and<br />

deduced that the recombination reaction would be first order<br />

in the density of TiO2(e )s at low light intensities, indicative of<br />

mechanism (24c), (24d) or (24e). 182,185 The former result has<br />

been previously observed 479 and the latter behavior was seen<br />

above <strong>with</strong> recombination of TiO 2(e )s and oxidized sensitizers<br />

in section 5/B/ii/b. 331,461 Using transient absorption spectroscopy<br />

and N3/TiO 2 thin-film electrodes in propylene<br />

carbonate electrolyte containing iodide, Durrant, Nelson,<br />

and co-workers monitored the kinetics for the loss of<br />

I2 . 382 After photo-excitation of B1% of the sensitizers,<br />

electron injection into TiO2, and hole <strong>transfer</strong> to iodide, I2<br />

was proposed to be formed by (11) or (10) followed by<br />

reaction (25). As evidenced by studies performed under variable-applied<br />

bias conditions, the rate of I2 loss was deter-<br />

mined to be second order in the concentration of I2 and 0th<br />

order in TiO2(e ) density consistent <strong>with</strong> the I2 dismutation<br />

reaction (18a). It was noted that the TiO2(e )s most likely<br />

recombined <strong>with</strong> I3 or I2 in a latter step but that neither<br />

species could be unambiguously identified in the spectra. Thus,<br />

this experiment modeled the fate of I2 but did not detail the<br />

reaction of the TiO2(e )s, which may be more relevant to the<br />

functioning DSSC. However, a mechanism consistent <strong>with</strong><br />

(24a) could be deduced.<br />

Durrant, Nelson, and co-workers studied the recombination<br />

kinetics for TiO 2(e )s <strong>with</strong> I 2 by transient absorption spectroscopy<br />

on unsensitized TiO 2 thin films in acetonitrile electrolytes<br />

by monitoring the loss of TiO 2(e )s after their formation<br />

by bandgap excitation and subsequent sacrificial hole scavenging<br />

<strong>with</strong> MeOH or Fe(CN)6 4 . 463 By transiently generating<br />

o13 TiO2(e )s/particle, it was shown that in the limit of a high<br />

concentration of I2, the TiO2(e ) kinetics followed the multiple-trapping,<br />

nearest-neighbor CTRW model and KWW function<br />

observed for trap-limited recombination. In the limit of a<br />

low concentration of I2 the kinetics became mono-exponential,<br />

as predicted by the dispersive, electron-<strong>transfer</strong> theory when<br />

<strong>heterogeneous</strong> electron <strong>transfer</strong> is limited by the concentration<br />

of acceptors, Fig. 45. Based on Monte Carlo simulations<br />

performed to mimic this behavior and a plot of half-lives<br />

versus the concentration of I2, second-order processes were<br />

ruled out. The kinetics for the loss of TiO2(e )s in the presence<br />

of 50 mM I2 was found to be over two orders-of-magnitude<br />

faster in the absence of 0.7 M LiI, and more dispersive. Given<br />

the strongly favorable equilibrium of I2 + I - I3 in<br />

acetonitrile, Keq 4 10 6 M 1 (eqn (9)), 344–349 this implies that<br />

I2 is a better acceptor than I3 for TiO2(e )s in acetonitrile.<br />

Fig. 45 Time-resolved, single-wavelength absorption difference spectra<br />

for unsensitized TiO2 thin films as a function of I2 concentration.<br />

Inset: A log–log plot of the half-life versus the concentration of I 2<br />

illustrating the proposed first-order recombination behavior in the<br />

concentration of I 2. Taken from Fig. 5 of ref. 463.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 157


The lack of any signal for I2 during recombination implied<br />

that the reaction mechanism for the loss of TiO2(e )s to the<br />

I 3 /I redox mediator on unsensitized TiO 2 thin films is<br />

analogous to the same reaction at the platinum counter<br />

electrode, i.e. dissociative adsorption followed by electron<br />

<strong>transfer</strong> and mechanism (24d).<br />

6. Conclusions<br />

It has been eighteen years since the celebrated Grätzel and<br />

O’Regan paper first appeared in Nature. This review demonstrates<br />

the tremendous progress that has been made towards<br />

developing a molecular-level understanding of <strong>charge</strong>-<strong>transfer</strong><br />

processes at sensitized TiO2 interfaces. The time scales and<br />

dynamics for excited-state electron injection into TiO2 have<br />

been quantified precisely under many experimental conditions.<br />

Regeneration of the photo-oxidized sensitizer by a variety of<br />

outer-sphere electron donors, including iodide, has also been<br />

quantified in some detail. Much less progress has been made<br />

towards our understanding of the unwanted <strong>charge</strong> recombination<br />

to oxidized iodide species, i.e. TiO 2(e ) + A. This is at least<br />

in part due to the inefficiency of these processes which makes<br />

characterization difficult. Fundamental data on the identity of<br />

the Acceptor(s) as well as the reduction mechanism(s) are still<br />

lacking. Nevertheless, studies have shown that the sensitizer p*<br />

orbitals and <strong>charge</strong>d ions in the electrolytes can play specific<br />

roles. Given the recent breakthroughs and the keen interest in<br />

these reactions, rapid progress is expected. A molecular-level<br />

understanding of the mechanisms for <strong>charge</strong> separation and<br />

recombination at sensitized semiconductor interfaces may ultimately<br />

enable optimal light-to-electrical power conversion in<br />

DSSCs and in future-generation photovoltaics.<br />

Acknowledgements<br />

The Division of Chemical Sciences, Office of Basic Energy<br />

Sciences, Office of Energy Research, U.S. Department of<br />

Energy, the National Science Foundation, and the donors of<br />

the Petroleum Research Fund, administered by the ACS, are<br />

gratefully acknowledged for research support.<br />

References<br />

1 M. I. Hoffert, K. Caldeira, A. K. Jain, E. F. Haites, L. D. D.<br />

Harvey, S. D. Potter, M. E. Schlesinger, S. H. Schneider,<br />

R. G. Watts, T. M. L. Wigley and D. J. Wuebbles, Nature,<br />

1998, 395, 881–884.<br />

2 K. Caldeira, A. K. Jain and M. I. Hoffert, Science, 2003, 299,<br />

2052–2054.<br />

3 United States, Department of Energy, Energy Information<br />

Administration, http://www.eia.doe.gov/.<br />

4 N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A.,<br />

2006, 103, 15729–15735.<br />

5D.Lu¨thi, M. Le Floch, B. Bereiter, T. Blunier, J.-M. Barnola,<br />

U. Siegenthaler, D. Raynaud, J. Jouzel, H. Fischer,<br />

K. Kawamura and T. F. Stocker, Nature, 2008, 453, 379–382.<br />

6 L. Loulergue, A. Schilt, R. Spahni, V. Masson-Delmotte,<br />

T. Blunier, B. Lemieux, J.-M. Barnola, D. Raynaud,<br />

T. F. Stocker and J. Chappellaz, Nature, 2008, 453, 383–386.<br />

7 Lewis, Nathan S., Powering the Planet—Global Energy Perspective,<br />

http://nsl.caltech.edu/energy.html.<br />

8 Department of Energy, Report of the Basic Energy Sciences<br />

Workshop on Solar Energy Utilization, Washington, DC, 2005.<br />

9 United Nations, World Energy Assessment Report: Energy and<br />

the Challenge of Sustainability, New York, NY, 2003.<br />

10 B. O’Regan and M. Gra¨tzel, Nature, 1991, 353, 737–740.<br />

11 A. Hagfeldt and M. Grätzel, Chem. Rev., 1995, 95, 49–68.<br />

12 M. A. Green, K. Emery, Y. Hishikawa and W. Warta, Progress in<br />

Photovoltaics: Research and Applications, 2008, 16, 61–67.<br />

13 A. Hagfeldt and M. Grätzel, Acc. Chem. Res., 2000, 33, 269–277.<br />

14 M. Grätzel, in a talk at the 12th International Conference on<br />

Photochemical Conversion and Storage of Solar Energy, Berlin,<br />

Germany, August, 1998.<br />

15 N. Papageorgiou, Coord. Chem. Rev., 2004, 248, 1421–1446.<br />

16 J. van de Lagemaat, K. Zhu, K. D. Benkstein and A. J. Frank,<br />

Inorg. Chim. Acta, 2008, 361, 620–626.<br />

17 A. J. Frank, N. Kopidakis and J. v. d. Lagemaat, Coord. Chem.<br />

Rev., 2004, 248, 1165–1179.<br />

18 L. M. Peter, Phys. Chem. Chem. Phys., 2007, 9, 2630–2642.<br />

19 J. Bisquert, Phys. Chem. Chem. Phys., 2008, 10, 3175–3194.<br />

20 J. Bisquert, Phys. Chem. Chem. Phys., 2008, 10, 49–72.<br />

21 National Renewable Energy Laboratory, Reference Solar Spectral<br />

Irradiance: Air Mass 1.5, http://rredc.nrel.gov/solar/spectra/am1.5/.<br />

22 K. Laqua, W. H. Melhuish and M. Zander, Pure Appl. Chem.,<br />

1988, 60, 1449–1460.<br />

23 G. M. Hasselman, D. F. Watson, J. R. Stromberg, D. F. Bocian,<br />

D. Holten, J. S. Lindsey and G. J. Meyer, J. Phys. Chem. B, 2006,<br />

110, 25430–25440.<br />

24 G. P. Smestad and M. Gra¨tzel, J. Chem. Educ., 1998, 75,<br />

752–756.<br />

25 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,<br />

E. Mueller, P. Liska, N. Vlachopoulos and M. Grätzel, J. Am.<br />

Chem. Soc., 1993, 115, 6382–6390.<br />

26 M. Gra¨tzel, J. Photochem. Photobiol. C: Photochem. Rev., 2003,<br />

4, 145–153.<br />

27 J. P. Paris and W. W. Brandt, J. Am. Chem. Soc., 1959, 81,<br />

5001–5002.<br />

28 F. E. Lytle and D. M. Hercules, J. Am. Chem. Soc., 1969, 91,<br />

253–257.<br />

29 F. Felix, J. Ferguson, H. U. Guedel and A. Ludi, J. Am. Chem.<br />

Soc., 1980, 102, 4096–4102.<br />

30 G. A. Crosby, K. W. Hipps and W. H. Elfring, J. Am. Chem. Soc.,<br />

1974, 96, 629–630.<br />

31 E. M. Kober and T. J. Meyer, Inorg. Chem., 1982, 21, 3967–3977.<br />

32 E. U. Condon, Phys. Rev., 1928, 32, 858–872.<br />

33 E. Condon, Phys. Rev., 1926, 28, 1182–1201.<br />

34 J. Franck and E. G. Dymond, Trans. Faraday Soc., 1926, 21,<br />

536–542.<br />

35 S. Wallin, J. Davidsson, J. Modin and L. Hammarstrom, J. Phys.<br />

Chem. A, 2005, 109, 4697–4704.<br />

36 A. T. Yeh, C. V. Shank and J. K. McCusker, Science, 2000, 289,<br />

935–938.<br />

37 E. M. Kober, B. P. Sullivan and T. J. Meyer, Inorg. Chem., 1984,<br />

23, 2098–2104.<br />

38 R. F. Dallinger and W. H. Woodruff, J. Am. Chem. Soc., 1979,<br />

101, 4391–4393.<br />

39 D. H. Oh and S. G. Boxer, J. Am. Chem. Soc., 1989, 111,<br />

1130–1131.<br />

40 F. Alary, J. L. Heully, L. Bijeire and P. Vicendo, Inorg. Chem.,<br />

2007, 46, 3154–3165.<br />

41 A. W. Adamson and J. N. Demas, J. Am. Chem. Soc., 1971, 93,<br />

1800–1801.<br />

42 G. A. Crosby and J. N. Demas, J. Am. Chem. Soc., 1971, 93,<br />

2841–2847.<br />

43 J. N. Demas and D. G. Taylor, Inorg. Chem., 1979, 18,<br />

3177–3179.<br />

44 H. Yersin and E. Gallhuber, J. Am. Chem. Soc., 1984, 106,<br />

6582–6586.<br />

45 H. Yersin, E. Gallhuber, A. Vogler and H. Kunkely, J. Am.<br />

Chem. Soc., 1983, 105, 4155–4156.<br />

46 G. D. Hager and G. A. Crosby, J. Am. Chem. Soc., 1975, 97,<br />

7031–7037.<br />

47 G. D. Hager, R. J. Watts and G. A. Crosby, J. Am. Chem. Soc.,<br />

1975, 97, 7037–7042.<br />

48 R. W. Harrigan and G. A. Crosby, J. Chem. Phys., 1973, 59,<br />

3468–3476.<br />

49 R. W. Harrigan, G. D. Hager and G. A. Crosby, Chem. Phys.<br />

Lett., 1973, 21, 487–490.<br />

158 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


50 K. W. Hipps and G. A. Crosby, J. Am. Chem. Soc., 1975, 97,<br />

7042–7048.<br />

51 P. D. Fleischauer, A. W. Adamson and G. Sartori, Prog. Inorg.<br />

Chem., 1972, 17, 1–56.<br />

52 A. W. Adamson, J. Chem. Educ., 1983, 60, 797–802.<br />

53 C. Daul, E. J. Baerends and P. Vernooijs, J. Am. Chem. Soc.,<br />

1994, 33, 3538–3543.<br />

54 P. Qu, D. W. Thompson and G. J. Meyer, Langmuir, 2000, 16,<br />

4662–4671.<br />

55 C. J. Timpson, C. A. Bignozzi, B. P. Sullivan, E. M. Kober and<br />

T. J. Meyer, J. Phys. Chem., 1996, 100, 2915–2925.<br />

56 P. Persson, R. Bergström, L. Ojamäe and S. Lunell, Quantum-<br />

Chemical Studies of Metal Oxides for Photoelectrochemical<br />

Applications, Adv. Quantum Chem., 2002, 41, 203–263.<br />

57 H. Rensmo, S. Lunell and H. Siegbahn, J. Photochem. Photobiol.<br />

A: Chem., 1998, 114, 117–124.<br />

58 H. Rensmo, S. So¨dergren, L. Patthey, K. Westermark,<br />

L. Vayssieres, O. Kohle, P. A. Bru¨hwiler, A. Hagfeldt and<br />

H. Siegbahn, Chem. Phys. Lett., 1997, 274, 51–57.<br />

59 A. J. Bard, G. M. Whitesides, R. N. Zare and F. W. McLafferty,<br />

Acc. Chem. Res., 1995, 28, 91–91.<br />

60 A. J. Bard and M. A. Fox, Acc. Chem. Res., 1995, 28, 141–145.<br />

61 M. K. Nazeeruddin, S. M. Zakeeruddin, R. Humphry-Baker,<br />

M. Jirousek, P. Liska, N. Vlachopoulos, V. Shklover,<br />

C. H. Fischer and M. Grätzel, Inorg. Chem., 1999, 38, 6298–6305.<br />

62 S. M. Zakeeruddin, M. K. Nazeeruddin, R. Humphry-Baker,<br />

P. Pechy, P. Quagliotto, C. Barolo, G. Viscardi and M. Grätzel,<br />

Langmuir, 2002, 18, 952–954.<br />

63 P. Wang, S. M. Zakeeruddin, J. E. Moser, M. K. Nazeeruddin,<br />

T. Sekiguchi and M. Grätzel, Nat. Mater., 2003, 2, 402–407.<br />

64 M. K. Nazeeruddin, S. M. Zakeeruddin, J. J. Lagref, P. Liska,<br />

P. Comte, C. Barolo, G. Viscardi, K. Schenk and M. Grätzel,<br />

Coord. Chem. Rev., 2004, 248, 1317–1328.<br />

65 P. Wang, C. Klein, R. Humphy-Baker, S. M. Zakeeruddin and<br />

M. Grätzel, Appl. Phys. Lett., 2005, 86, 123508–123510.<br />

66 P. Wang, C. Klein, R. Humphry-Baker, S. M. Zakeeruddin and<br />

M. Grätzel, J. Am. Chem. Soc., 2005, 127, 808–809.<br />

67 D. Kuang, S. Ito, B. Wenger, C. Klein, J. E. Moser, R. Humphry-<br />

Baker, S. M. Zakeeruddin and M. Gra¨tzel, J. Am. Chem. Soc.,<br />

2006, 128, 4146–4154.<br />

68 M. K. Nazeeruddin, P. Pe´chy and M. Grätzel, Chem. Commun.,<br />

1997, 1705–1706.<br />

69 C. R. Bock, J. A. Connor, A. R. Gutierrez, T. J. Meyer,<br />

D. G. Whitten, B. P. Sullivan and J. K. Nagle, J. Am. Chem.<br />

Soc., 1979, 101, 4815–4824.<br />

70 C. R. Bock, T. J. Meyer and D. G. Whitten, J. Am. Chem. Soc.,<br />

1975, 97, 2909–2911.<br />

71 P. Bonhote, E. Gogniat, S. Tingry, C. Barbe, N. Vlachopoulos,<br />

F. Lenzmann, P. Comte and M. Gra¨tzel, J. Phys. Chem. B, 1998,<br />

102, 1498–1507.<br />

72 T. A. Heimer, S. T. D’Arcangelis, F. Farzad, J. M. Stipkala and<br />

G. J. Meyer, Inorg. Chem., 1996, 35, 5319–5324.<br />

73 S. A. Trammell and T. J. Meyer, J. Phys. Chem. B, 1999, 103, 104–107.<br />

74 D. Rehm and A. Weller, Ber. Bunsen-Ges.Phys. Chem., 1969, 73,<br />

834–839.<br />

75 D. Rehm and A. Weller, Isr. J. Chem., 1970, 8, 259–271.<br />

76 P. Chen and T. J. Meyer, Chem. Rev., 1998, 98, 1439–1478.<br />

77 M. Yang, D. W. Thompson and G. J. Meyer, Inorg. Chem., 2002,<br />

41, 1254–1262.<br />

78 R. Argazzi, C. A. Bignozzi, M. Yang, G. M. Hasselmann and<br />

G. J. Meyer, Nano Lett., 2002, 2, 625–628.<br />

79 L. C. T. Shoute and G. R. Loppnow, J. Am. Chem. Soc., 2003,<br />

125, 15636–15646.<br />

80 K. Kalyanasundaram, N. Vlachopoulos, V. Krishnan,<br />

A. Monnier and M. Grätzel, J. Phys. Chem., 1987, 91, 2342–2347.<br />

81 A. Staniszewski, W. B. Heuer and G. J. Meyer, Inorg. Chem.,<br />

2008, 47, 7062–7064.<br />

82 H. J. Snaith, C. S. Karthikeyan, A. Petrozza, J. Teuscher,<br />

J. E. Moser, M. K. Nazeeruddin, M. Thelakkat and<br />

M. Grätzel, J. Phys. Chem. C, 2008, 112, 7562–7566.<br />

83 V. Aranyos, J. Hjelm, A. Hagfeldt and H. Grennberg, J. Chem.<br />

Soc., Dalton Trans., 2001, 1319–1325.<br />

84 P. Wang, S. M. Zakeeruddin, J. E. Moser, R. Humphry-Baker,<br />

P. Comte, V. Aranyos, A. Hagfeldt, M. K. Nazeeruddin and<br />

M. Grätzel, Adv. Mater., 2004, 16, 1806–1811.<br />

85 S. R. Jang, C. Lee, H. Choi, J. J. Ko, J. Lee, R. Vittal<br />

and K. J. Kim, Chem. Mater., 2006, 18, 5604–5608.<br />

86 Y. j. Hou, P. h. Xie, B. w. Zhang, Y. Cao, X. r. Xiao and<br />

W. b. Wang, Inorg. Chem., 1999, 38, 6320–6322.<br />

87 F. Gao, Y. Wang, J. Zhang, D. Shi, M. Wang, R. Humphry-<br />

Baker, P. Wang, S. M. Zakeeruddin and M. Gra¨tzel, Chem.<br />

Commun., 2008, 2635–2637.<br />

88 V. Balzani, L. Moggi and F. Scandola, in: Towards a Supramolecular<br />

Photochemistry: Assembly of Molecular Components to<br />

Obtain Photochemical Molecular Devices, ed. V. Balzani,<br />

Dordrecht, Holland, 1987.<br />

89 C. A. Bignozzi, R. Argazzi, M. T. Indelli and F. Scandola, Sol.<br />

Energy Mater., 1994, 32, 229–244.<br />

90 C. A. Bignozzi, R. Argazzi, F. Scandola, J. R. Schoonover and<br />

G. J. Meyer, Sol. Energy Mater., 1995, 38, 187–198.<br />

91 R. Amadelli, R. Argazzi, C. A. Bignozzi and F. Scandola, J. Am.<br />

Chem. Soc., 1990, 112, 7099–7103.<br />

92 M. K. Nazeeruddin, P. Liska, J. Moser, N. Vlachopoulos and<br />

M. Gra¨tzel, Helv. Chim. Acta, 1990, 73, 1788–1803.<br />

93 D. Holten, D. F. Bocian and J. S. Lindsey, Acc. Chem. Res., 2002,<br />

35, 57–69.<br />

94 R. J. Forster, T. E. Keyes and J. G. Vos, Interfacial Supramolecular<br />

Assemblies, John Wiley & Sons Ltd., 2003.<br />

95 F. Gajardo, A. M. Leiva, B. Loeb, A. Delgadillo, J. R. Stromberg<br />

and G. J. Meyer, Inorg. Chim. Acta, 2008, 361, 613–619.<br />

96 J. Van Houten and R. J. Watts, J. Am. Chem. Soc., 1976, 98,<br />

4853–4858.<br />

97 W. Siebrand, J. Chem. Phys., 1966, 44, 4055–4057.<br />

98 R. Englman and J. Jortner, Mol. Phys., 1970, 18, 145–164.<br />

99 K. F. Freed and J. Jortner, J. Chem. Phys., 1970, 52, 6272–6291.<br />

100 M. Bixon and J. Jortner, J. Chem. Phys., 1968, 48, 715–726.<br />

101 G. W. Robinson and R. P. Frosch, J. Chem. Phys., 1963, 38,<br />

1187–1203.<br />

102 N. A. Anderson and T. Lian, Coord. Chem. Rev., 2004, 248,<br />

1231–1246.<br />

103 J. B. Asbury, N. A. Anderson, E. Hao, X. Ai and T. Lian, J. Phys.<br />

Chem. B, 2003, 107, 7376–7386.<br />

104 J. B. Asbury, E. Hao, Y. Wang, H. N. Ghosh and T. Lian,<br />

J. Phys. Chem. B, 2001, 105, 4545–4557.<br />

105 J. B. Asbury, E. Hao, Y. Wang and T. Lian, J. Phys. Chem. B,<br />

2000, 104, 11957–11964.<br />

106 J. B. Asbury, Y.-Q. Wang, E. Hao, H. N. Ghosh and T. Lian,<br />

Res. Chem. Intermed., 2001, 27, 393–406.<br />

107 G. Benkö, J. Kallioinen, J. E. I. Korppi-Tommola, A. P. Yartsev<br />

and V. Sundström, J. Am. Chem. Soc., 2002, 124, 489–493.<br />

108 R. Ernstorfer, L. Gundlach, S. Felber, W. Storck, R. Eichberger<br />

and F. Willig, J. Phys. Chem. B, 2006, 110, 25383–25391.<br />

109 T. Hannappel, B. Burfeindt, W. Storck and F. Willig, J. Phys.<br />

Chem. B, 1997, 101, 6799–6802.<br />

110 J. Kallioinen, G. Benko¨, P. Myllyperkio¨, L. Khriachtchev,<br />

B. Skarman, R. Wallenberg, M. Tuomikoski, J. Korppi-Tommola,<br />

V. Sundström and A. P. Yartsev, J. Phys. Chem. B, 2004, 108,<br />

6365–6373.<br />

111 J. Kallioinen, G. Benkö, V. Sundström, J. E. I. Korppi-Tommola<br />

and A. P. Yartsev, J. Phys. Chem. B, 2002, 106, 4396–4404.<br />

112 D. Kuciauskas, J. E. Monat, R. Villahermosa, H. B. Gray,<br />

N. S. Lewis and J. K. McCusker, J. Phys. Chem. B, 2002, 106,<br />

9347–9358.<br />

113 A. Morandeira, G. Boschloo, A. Hagfeldt and L. Hammarstro¨m,<br />

J. Phys. Chem. B, 2005, 109, 19403–19410.<br />

114 P. Myllyperkiö, G. Benkö, J. Korppi-Tommola, A. P. Yartsev<br />

and V. Sundström, Phys. Chem. Chem. Phys., 2008, 10, 996–1002.<br />

115 K. Schwarzburg, R. Ernstorfer, S. Felber and F. Willig, Coord.<br />

Chem. Rev., 2004, 248, 1259–1270.<br />

116 C. She, J. Guo, S. Irle, K. Morokuma, D. L. Mohler, H. Zabri,<br />

F. Odobel, K. T. Youm, F. Liu, J. T. Hupp and T. Lian, J. Phys.<br />

Chem. A, 2007, 111, 6832–6842.<br />

117 Y. Tachibana, S. A. Haque, I. P. Mercer, J. R. Durrant and<br />

D. R. Klug, J. Phys. Chem. B, 2000, 104, 1198–1205.<br />

118 Y. Tachibana, J. E. Moser, M. Grätzel, D. R. Klug and<br />

J. R. Durrant, J. Phys. Chem., 1996, 100, 20056–20062.<br />

119 D. F. Watson and G. J. Meyer, Annu. Rev. Phys. Chem., 2005, 56,<br />

119–156.<br />

120 M. A. Webb, F. J. Knorr and J. L. McHale, J. Raman Spectrosc.,<br />

2001, 32, 481–485.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 159


121 L. F. Cooley, P. Bergquist and D. F. Kelley, J. Am. Chem. Soc.,<br />

1990, 112, 2612–2617.<br />

122 R. A. Malone and D. F. Kelley, J. Chem. Phys., 1991, 95,<br />

8970–8976.<br />

123 A. C. Bhasikuttan, M. Suzuki, S. Nakashima and T. Okada,<br />

J. Am. Chem. Soc., 2002, 124, 8398–8405.<br />

124 S. Cazzanti, S. Caramori, R. Argazzi, C. M. Elliott and<br />

C. A. Bignozzi, J. Am. Chem. Soc., 2006, 128, 9996–9997.<br />

125 S. Yoon, P. Kukura, C. M. Stuart and R. A. Mathies, Mol. Phys.,<br />

2006, 104, 1275–1282.<br />

126 N. H. Damrauer, G. Cerullo, A. Yeh, T. R. Boussie, C. V. Shank<br />

and J. K. McCusker, Science, 1997, 275, 54–57.<br />

127 N. H. Damrauer and J. K. McCusker, J. Phys. Chem. A, 1999,<br />

103, 8440–8446.<br />

128 A. N. Tarnovsky, W. Gawelda, M. Johnson, C. Bressler and<br />

M. Chergui, J. Phys. Chem. B, 2006, 110, 26497–26505.<br />

129 W. Henry, C. G. Coates, C. Brady, K. L. Ronayne, P. Matousek,<br />

M. Towrie, S. W. Botchway, A. W. Parker, J. G. Vos,<br />

W. R. Browne and J. J. McGarvey, J. Phys. Chem. A, 2008,<br />

112, 4537–4544.<br />

130 W. D. K. Clark and N. Sutin, J. Am. Chem. Soc., 1977, 99,<br />

4676–4682.<br />

131 M. Gleria and R. Memming, Z. Phys. Chem. (Munich), 1975, 98,<br />

303–316.<br />

132 S. Anderson, E. C. Constable, M. P. Dare-Edwards,<br />

J. B. Goodenough, A. Hamnett, K. R. Seddon and<br />

R. D. Wright, Nature, 1979, 280, 571–573.<br />

133 A. Marton, C. C. Clark, R. Srinivasan, R. E. Freundlich,<br />

A. A. Narducci Sarjeant and G. J. Meyer, Inorg. Chem., 2006,<br />

45, 362–369.<br />

134 T. J. Meyer, G. J. Meyer, B. W. Pfennig, J. R. Schoonover,<br />

C. J. Timpson, J. F. Wall, C. Kobusch, X. Chen, B. M. Peek,<br />

C. G. Wall, W. Ou, B. W. Erickson and C. A. Bignozzi, Inorg.<br />

Chem., 1994, 33, 3952–3964.<br />

135 P. G. Hoertz, A. Staniszewski, A. Marton, G. T. Higgins,<br />

C. D. Incarvito, A. L. Rheingold and G. J. Meyer, J. Am. Chem.<br />

Soc., 2006, 128, 8234–8245.<br />

136 P. Persson, S. Lunell and L. Ojamäe, Chem. Phys. Lett., 2002,<br />

364, 469–474.<br />

137 P. Qu and G. J. Meyer, Langmuir, 2001, 17, 6720–6728.<br />

138 S. Umapathy, A. M. Cartner, A. W. Parker and R. E. Hester,<br />

J. Phys. Chem., 1990, 94, 8880–8885.<br />

139 K. S. Finnie, J. R. Bartlett and J. L. Woolfrey, Langmuir, 1998,<br />

14, 2744–2749.<br />

140 K. Kilsa, E. I. Mayo, B. S. Brunschwig, H. B. Gray, N. S.<br />

Lewis and J. R. Winkler, J. Phys. Chem. B, 2004, 108,<br />

15640–15651.<br />

141 K. D. Dobson and A. J. McQuillan, Spectrochim. Acta, Part A,<br />

2000, 56, 557–565.<br />

142 G. B. Deacon and R. J. Phillips, Coord. Chem. Rev., 1980, 33,<br />

227–250.<br />

143 P. Persson, R. Bergstrom and S. Lunell, J. Phys. Chem. B, 2000,<br />

104, 10348–10351.<br />

144 P. Pechy, F. P. Rotzinger, M. K. Nazeeruddin, O. Kohle, Shaik<br />

M. Zakeeruddin, R. Humphry-Baker and M. Gra¨tzel, J. Chem.<br />

Soc., Chem. Commun., 1995, 65–66.<br />

145 T. A. Heimer, C. A. Bignozzi and G. J. Meyer, J. Phys. Chem.,<br />

1993, 97, 11987–11994.<br />

146 G. J. Meyer, Inorg. Chem., 2005, 44, 6852–6864.<br />

147 T. Bessho, E. C. Constable, M. Gra¨tzel, A. H. Redondo,<br />

C. E. Housecroft, W. Kylberg, M. K. Nazeeruddin,<br />

M. Neuburger and S. Schaffner, Chem. Commun., 2008,<br />

3717–3719.<br />

148 S. Sakaki, T. Kuroki and T. Hamada, J. Chem. Soc., Dalton<br />

Trans., 2002, 840–842.<br />

149 N. Alonso-Vante, J.-F. Nierengarten and J.-P. Sauvage, J. Chem.<br />

Soc., Dalton Trans., 1994, 1649–1654.<br />

150 C. A. Kelly, F. Farzad, D. W. Thompson and G. J. Meyer,<br />

Langmuir, 1999, 15, 731–737.<br />

151 F. Farzad, D. W. Thompson, C. A. Kelly and G. J. Meyer, J. Am.<br />

Chem. Soc., 1999, 121, 5577–5578.<br />

152 G. T. Higgins, B. V. Bergeron, G. M. Hasselmann, F. Farzad and<br />

G. J. Meyer, J. Phys. Chem. B, 2006, 110, 2598–2605.<br />

153 H. Gerischer and F. Willig, Top. Curr. Chem., 1976, 61, 31–84.<br />

154 H. Gerischer, Photochem. Photobiol., 1972, 16, 243–260.<br />

155 H. Gerischer, Surf. Sci., 1969, 18, 97–122.<br />

156 K. Westermark, A. Henningsson, H. Rensmo, S. So¨dergren,<br />

H. Siegbahn and A. Hagfeldt, Chem. Phys., 2002, 285,<br />

157–165.<br />

157 F. Cardon and W. P. Gomes, J. Phys. D: Appl. Phys., 1978, 11,<br />

L63–L67.<br />

158 B. O’Regan, J. Moser, M. Anderson and M. Gra¨tzel, J. Phys.<br />

Chem., 1990, 94, 8720–8726.<br />

159 B. A. Gregg, Coord. Chem. Rev., 2004, 248, 1215–1224.<br />

160 K. Schwarzburg and F. Willig, J. Phys. Chem. B, 1999, 103,<br />

5743–5746.<br />

161 A. Zaban, A. Meier and B. A. Gregg, J. Phys. Chem. B, 1997, 101,<br />

7985–7990.<br />

162 J. Bisquert, G. Garcia-Belmonte and F. Fabregat-Santiago,<br />

J. Solid State Electrochem., 1999, 3, 337–347.<br />

163 W. J. Albery and P. N. Bartlett, J. Electrochem. Soc., 1984, 131,<br />

315–325.<br />

164 G. Hodes, I. D. J. Howell and L. M. Peter, J. Electrochem. Soc.,<br />

1992, 139, 3136–3140.<br />

165 A. Hagfeldt, U. Bjo¨rkstén and S.-E. Lindquist, Sol. Energy<br />

Mater., 1992, 27, 293–304.<br />

166 G. Rothenberger, D. Fitzmaurice and M. Grätzel, J. Phys.<br />

Chem., 1992, 96, 5983–5986.<br />

167 B. Enright, G. Redmond and D. Fitzmaurice, J. Phys. Chem.,<br />

1994, 98, 6195–6200.<br />

168 G. Redmond and D. Fitzmaurice, J. Phys. Chem., 1993, 97,<br />

1426–1430.<br />

169 L. A. Lyon and J. T. Hupp, J. Phys. Chem. B, 1999, 103,<br />

4623–4628.<br />

170 D. Fitzmaurice, Sol. Energy Mater., 1994, 32, 289–305.<br />

171 J. Bisquert, F. Fabregat-Santiago, I. Mora-Sero, G. Garcia-Belmonte,<br />

E. M. Barea and E. Palomares, Inorg. Chim. Acta, 2008,<br />

361, 684–698.<br />

172 V. G. Kytin, J. Bisquert, I. Abayev and A. Zaban, Phys. Rev. B,<br />

2004, 70, 193304.<br />

173 M. Bailes, P. J. Cameron, K. Lobato and L. M. Peter, J. Phys.<br />

Chem. B, 2005, 109, 15429–15435.<br />

174 I. Abayev, A. Zaban, V. G. Kytin, A. A. Danilin, G. Garcia-<br />

Belmonte and J. Bisquert, J. Solid State Electrochem., 2007, 11,<br />

647–653.<br />

175 F. Fabregat-Santiago, I. Mora-Sero, G. Garcia-Belmonte and<br />

J. Bisquert, J. Phys. Chem. B, 2003, 107, 758–768.<br />

176 A. J. Morris and G. J. Meyer, J. Phys. Chem. C, 2008, 112,<br />

18224–18231.<br />

177 S. A. Haque, E. Palomares, B. M. Cho, A. N. M. Green,<br />

N. Hirata, D. R. Klug and J. R. Durrant, J. Am. Chem. Soc.,<br />

2005, 127, 3456–3462.<br />

178 J. R. Durrant, J. Photochem. Photobiol. A: Chem., 2002, 148,<br />

5–10.<br />

179 Y. Tachibana, I. V. Rubtsov, I. Montanari, K. Yoshihara,<br />

D. R. Klug and J. R. Durrant, J. Photochem. Photobiol. A:<br />

Chem., 2001, 142, 215–220.<br />

180 J. van de Lagemaat, N. Kopidakis, N. R. Neale and A. J. Frank,<br />

Phys. Rev. B, 2005, 71, 035304.<br />

181 P. E. de Jongh and D. Vanmaekelbergh, Phys. Rev. Lett., 1996,<br />

77, 3427–3430.<br />

182 J. R. Durrant, S. A. Haque and E. Palomares, Coord. Chem. Rev.,<br />

2004, 248, 1247–1257.<br />

183 J. Nelson, Phys. Rev. B, 1999, 59, 15374.<br />

184 J. Nelson and R. E. Chandler, Coord. Chem. Rev., 2004, 248,<br />

1181–1194.<br />

185 J. Nelson, S. A. Haque, D. R. Klug and J. R. Durrant, Phys. Rev.<br />

B, 2001, 63, 205321.<br />

186 A. B. Walker, L. M. Peter, D. Martínez and K. Lobato, Chimia,<br />

2007, 61, 792–795.<br />

187 G. Boschloo and D. Fitzmaurice, J. Phys. Chem. B, 1999, 103,<br />

2228–2231.<br />

188 J. Moser, S. Punchihewa, P. P. Infelta and M. Grätzel, Langmuir,<br />

1991, 7, 3012–3018.<br />

189 G. Redmond, D. Fitzmaurice and M. Grätzel, J. Phys. Chem.,<br />

1993, 97, 6951–6954.<br />

190 L. Kavan, K. Kratochvilova´ and M. Gra¨tzel, J. Electroanal.<br />

Chem., 1995, 394, 93–102.<br />

191 H. Wang, J. He, G. Boschloo, H. Lindstrom, A. Hagfeldt and<br />

S. E. Lindquist, J. Phys. Chem. B, 2001, 105, 2529–2533.<br />

160 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


192 L. de la Garza, Z. V. Saponjic, N. M. Dimitrijevic,<br />

M. C. Thurnauer and T. Rajh, J. Phys. Chem. B, 2006, 110,<br />

680–686.<br />

193 T. Berger, T. Lana-Villarreal, D. Monllor-Satoca and R. Go´mez,<br />

Electrochem. Commun., 2006, 8, 1713–1718.<br />

194 G. Ramakrishna, A. Das and H. N. Ghosh, Langmuir, 2004, 20,<br />

1430–1435.<br />

195 T. Dittrich, Physica Status Solidi A, 2000, 182, 447–455.<br />

196 F. Urbach, Phys. Rev., 1953, 92, 1324.<br />

197 H. Tang, F. Lévy, H. Berger and P. E. Schmid, Phys. Rev. B,<br />

1995, 52, 7771.<br />

198 V. Duzhko, V. Y. Timoshenko, F. Koch and T. Dittrich, Phys.<br />

Rev. B, 2001, 64, 075204.<br />

199 E. Burstein, Phys. Rev., 1954, 93, 632–633.<br />

200 T. S. Moss, Proc. Phys. Soc. B, 1954, 67, 775–782.<br />

201 B. O’Regan, M. Grätzel and D. Fitzmaurice, J. Phys. Chem.,<br />

1991, 95, 10525–10528.<br />

202 B. O’Regan, M. Gra¨tzel and D. Fitzmaurice, Chem. Phys. Lett.,<br />

1991, 183, 89–93.<br />

203 K. Takeshita, Y. Sasaki, M. Kobashi, Y. Tanaka, S. Maeda,<br />

A. Yamakata, T.-a. Ishibashi and H. Onishi, J. Phys. Chem. B,<br />

2004, 108, 2963–2969.<br />

204 A. Kay, R. Humphry-Baker and M. Gra¨tzel, J. Phys. Chem.,<br />

1994, 98, 952–959.<br />

205 G. Boschloo and D. Fitzmaurice, J. Phys. Chem. B, 1999, 103,<br />

7860–7868.<br />

206 A. Von Hippel, J. Kalnajs and W. B. Westphal, J. Phys. Chem.<br />

Solids, 1962, 23, 779–796.<br />

207 H. Lindstrom, S. Sodergren, A. Solbrand, H. Rensmo, J. Hjelm,<br />

A. Hagfeldt and S. E. Lindquist, J. Phys. Chem. B, 1997, 101,<br />

7710–7716.<br />

208 H. Lindstrom, S. Sodergren, A. Solbrand, H. Rensmo, J. Hjelm,<br />

A. Hagfeldt and S. E. Lindquist, J. Phys. Chem. B, 1997, 101,<br />

7717–7722.<br />

209 R. van de Krol, A. Goossens and E. A. Meulenkamp, J. Appl.<br />

Phys., 2001, 90, 2235–2242.<br />

210 R. van de Krol, A. Goossens and J. Schoonman, J. Phys. Chem.<br />

B, 1999, 103, 7151–7159.<br />

211 R. J. Cava, D. W. Murphy, S. Zahurak, A. Santoro and<br />

R. S. Roth, J. Solid State Chem., 1984, 53, 64–75.<br />

212 L. Kavan, M. Gra¨tzel, J. Rathousky and A. Zukalb,<br />

J. Electrochem. Soc., 1996, 143, 394–400.<br />

213 B. Zachau-Christiansen, K. West, T. Jacobsen and S. Atlung,<br />

Solid State Ionics, 1988, 28–30, 1176–1182.<br />

214 R. van de Krol, A. Goossens and E. A. Meulenkamp,<br />

J. Electrochem. Soc., 1999, 146, 3150–3154.<br />

215 A. Henningsson, M. P. Andersson, P. Uvdal, H. Siegbahn and<br />

A. Sandell, Chem. Phys. Lett., 2002, 360, 85–90.<br />

216 I. Exnar, L. Kavan, S. Y. Huang and M. Grätzel, J. Power<br />

Sources, 1997, 68, 720–722.<br />

217 S. Y. Huang, L. Kavan, I. Exnar and M. Grätzel, J. Electrochem.<br />

Soc., 1995, 142, L142–L144.<br />

218 J. I. Pankove, Optical Processes in Semiconductors, Courier Dover<br />

Publications, 1975.<br />

219 S. M. Sze and K. K. Ng, Physics of Semiconductor Devices, Wiley-<br />

Interscience, 2006.<br />

220 T. Berger, M. Sterrer, O. Diwald, E. Knozinger, D. Panayotov,<br />

T. L. Thompson and J. T. Yates, J. Phys. Chem. B, 2005, 109,<br />

6061–6068.<br />

221 F. Cao, G. Oskam, P. C. Searson, J. M. Stipkala, T. A. Heimer,<br />

F. Farzad and G. J. Meyer, J. Phys. Chem., 1995, 99,<br />

11974–11980.<br />

222 R. F. Howe and M. Gra¨tzel, J. Phys. Chem., 1985, 89, 4495–4499.<br />

223 A. C. Fisher, L. M. Peter, E. A. Ponomarev, A. B. Walker and K.<br />

G. U. Wijayantha, J. Phys. Chem. B, 2000, 104, 949–958.<br />

224 R. L. Willis, C. Olson, B. O’Regan, T. Lutz, J. Nelson and<br />

J. R. Durrant, J. Phys. Chem. B, 2002, 106, 7605–7613.<br />

225 C. Kittel, Introduction to Solid State Physics, John Wiley & Sons,<br />

Inc., 2004.<br />

226 A. Yamakata, T.-a. Ishibashi and H. Onishi, Chem. Phys. Lett.,<br />

2001, 333, 271–277.<br />

227 J. M. Bolts and M. S. Wrighton, J. Phys. Chem., 1976, 80,<br />

2641–2645.<br />

228 H. Gerischer, Semiconductor Electrochemistry, in Treatise on<br />

Physical Chemistry Vol. IXA: Electrochemistry, ed. H. Eyring,<br />

J. Henderson and W. Jost, Academic Press, New York, 1970,<br />

pp. 463–542.<br />

229 A. J. Bard and L. R. Faulkner, Electrochemical Methods: Fundamentals<br />

and Applications, John Wiley & Sons, Inc., 2001.<br />

230 Y. G. Berube and P. L. de Bruyn, J. Colloid Interface Sci., 1968,<br />

28, 92–105.<br />

231 K. Bourikas, T. Hiemstra and W. H. Van Riemsdijk, Langmuir,<br />

2001, 17, 749–756.<br />

232 M. Kosmulski and J. B. Rosenholm, J. Phys. Chem., 1996, 100,<br />

11681–11687.<br />

233 D. E. Yates and T. W. Healy, J. Chem. Soc., Faraday Trans. 1,<br />

1980, 76, 9–18.<br />

234 A. Loupy and B. Tchoubar, Salt Effects in Organic and Organo<strong>metal</strong>lic<br />

Chemistry, VCH Verlagsgesellschaft mbH, Weinheim,<br />

Germany, 1992.<br />

235 M. G. Evans and M. Polanyi, Trans. Faraday Soc., 1935, 31,<br />

875–894.<br />

236 H. Eyring, J. Chem. Phys., 1935, 3, 107–115.<br />

237 N. S. Hush, J. Chem. Phys., 1958, 28, 962–972.<br />

238 N. S. Hush, Trans. Faraday Soc., 1961, 57, 557–580.<br />

239 R. A. Marcus, Annu. Rev. Phys. Chem., 1964, 15, 155–196.<br />

240 R. A. Marcus and N. Sutin, Biochim. Biophys. Acta (BBA)—Rev.<br />

Bioenergetics, 1985, 811, 265–322.<br />

241 S. Ramakrishna, F. Willig, V. May and A. Knorr, J. Phys. Chem.<br />

B, 2003, 107, 607–611.<br />

242 W. Stier and O. V. Prezhdo, J. Phys. Chem. B, 2002, 106,<br />

8047–8054.<br />

243 F. Willig, C. Zimmermann, S. Ramakrishna and W. Storck,<br />

Electrochim. Acta, 2000, 45, 4565–4575.<br />

244 C. Zimmermann, F. Willig, S. Ramakrishna, B. Burfeindt,<br />

B. Pettinger, R. Eichberger and W. Storck, J. Phys. Chem. B,<br />

2001, 105, 9245–9253.<br />

245 Y. Tachibana, M. K. Nazeeruddin, M. Grätzel, D. R. Klug and<br />

J. R. Durrant, Chem. Phys., 2002, 285, 127–132.<br />

246 B. Wenger, M. Grätzel and J.-E. Moser, Chimia, 2005, 59, 123–125.<br />

247 B. Wenger, M. Gra¨tzel and J. E. Moser, J. Am. Chem. Soc., 2005,<br />

127, 12150–12151.<br />

248 T. D. M. Bell, C. Pagba, M. Myahkostupov, J. Hofkens and<br />

P. Piotrowiak, J. Phys. Chem. B, 2006, 110, 25314–25321.<br />

249 C. She, J. Guo and T. Lian, J. Phys. Chem. B, 2007, 111,<br />

6903–6912.<br />

250 S. Iwai, K. Hara, S. Murata, R. Katoh, H. Sugihara and<br />

H. Arakawa, J. Chem. Phys., 2000, 113, 3366–3373.<br />

251 G. Benko¨, B.Skarman,R.Wallenberg,A.Hagfeldt,V.Sundstro¨m<br />

andA.P.Yartsev,J. Phys. Chem. B, 2003, 107, 1370–1375.<br />

252 H. Nishikiori, W. Qian, M. A. El-Sayed, N. Tanaka and T. Fujii,<br />

J. Phys. Chem. C, 2007, 111, 9008–9011.<br />

253 E. Palomares, M. V. Martinez-Diaz, S. A. Haque, T. Torres and<br />

J. R. Durrant, Chem. Commun., 2004, 2112–2113.<br />

254 C. A. Kelly, F. Farzad, D. W. Thompson, J. M. Stipkala and<br />

G. J. Meyer, Langmuir, 1999, 15, 7047–7054.<br />

255 P. A. Connor, K. D. Dobson and A. J. McQuillan, Langmuir,<br />

1999, 15, 2402–2408.<br />

256 K. D. Dobson, P. A. Connor and A. J. McQuillan, Langmuir,<br />

1997, 13, 2614–2616.<br />

257 S. Ardizzone and S. Trasatti, Adv. Colloid Interface Sci., 1996, 64,<br />

173–251.<br />

258 M. Harju, M. Järn, P. Dahlsten, J. B. Rosenholm and<br />

T. Mäntylä, Appl. Surf. Sci., 2008, 254, 7272–7279.<br />

259 J. Ferguson, A. W. H. Mau and W. H. F. Sasse, Chem. Phys.<br />

Lett., 1979, 68, 21–24.<br />

260 P. V. Kamat, I. Bedja, S. Hotchandani and L. K. Patterson,<br />

J. Phys. Chem., 1996, 100, 4900–4908.<br />

261 B. I. Lemon and J. T. Hupp, J. Phys. Chem. B, 1999, 103,<br />

3797–3799.<br />

262 Y. Tachibana, S. A. Haque, I. P. Mercer, J. E. Moser, D. R. Klug<br />

and J. R. Durrant, J. Phys. Chem. B, 2001, 105, 7424–7431.<br />

263 G. L. Closs and J. R. Miller, Science, 1988, 240, 440–447.<br />

264 J. F. Smalley, H. O. Finklea, C. E. D. Chidsey, M. R. Linford,<br />

S. E. Creager, J. P. Ferraris, K. Chalfant, T. Zawodzinsk,<br />

S. W. Feldberg and M. D. Newton, J. Am. Chem. Soc., 2003,<br />

125, 2004–2013.<br />

265 J. F. Smalley, S. W. Feldberg, C. E. D. Chidsey, M. R. Linford,<br />

M. D. Newton and Y.-P. Liu, J. Phys. Chem., 1995, 99,<br />

13141–13149.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 161


266 H. B. Gray and J. R. Winkler, Proc. Natl. Acad. Sci. U. S. A.,<br />

2005, 102, 3534–3539.<br />

267 K. Kilsa, E. I. Mayo, D. Kuciauskas, R. Villahermosa,<br />

N. S. Lewis, J. R. Winkler and H. B. Gray, J. Phys. Chem. A,<br />

2003, 107, 3379–3383.<br />

268 E. Galoppini, Coord. Chem. Rev., 2004, 248, 1283–1297.<br />

269 E. Galoppini, W. Guo, P. Qu and G. J. Meyer, J. Am. Chem. Soc.,<br />

2001, 123, 4342–4343.<br />

270 P. Piotrowiak, E. Galoppini, Q. Wei, G. J. Meyer and P. Wiewior,<br />

J. Am. Chem. Soc., 2003, 125, 5278–5279.<br />

271 B. Wenger, C. Bauer, M. K. Nazeeruddin, P. Comte,<br />

S. M. Zakeeruddin, M. Grätzel and J.-E. Moser, Physical Chemistry<br />

of Interfaces and Nanomaterials V, ed. M. Spitler and<br />

F. Willig, Proceedings of SPIE, 2006, vol. 6325, p. 63250V.<br />

272 E. Palomares, J. N. Clifford, S. A. Haque, T. Lutz and<br />

J. R. Durrant, J. Am. Chem. Soc., 2003, 125, 475–482.<br />

273 A. R. Kortan, R. Hull, R. L. Opila, M. G. Bawendi,<br />

M. L. Steigerwald, P. J. Carroll and L. E. Brus, J. Am. Chem.<br />

Soc., 1990, 112, 1327–1332.<br />

274 I. Bedja and P. V. Kamat, J. Phys. Chem., 1995, 99, 9182–9188.<br />

275 Y. Diamant, S. Chappel, S. G. Chen, O. Melamed and A. Zaban,<br />

Coord. Chem. Rev., 2004, 248, 1271–1276.<br />

276 A. Zaban, S. G. Chen, S. Chappel and B. A. Gregg, Chem.<br />

Commun., 2000, 2231–2232.<br />

277 W. H. Rippard, A. C. Perrella, F. J. Albert and R. A. Buhrman,<br />

Phys. Rev. Lett., 2002, 88, 046805.<br />

278 M. Myahkostupov, P. Piotrowiak, D. Wang and E. Galoppini,<br />

J. Phys. Chem. C, 2007, 111, 2827–2829.<br />

279 N. A. Anderson, X. Ai, D. Chen, D. L. Mohler and T. Lian,<br />

J. Phys. Chem. B, 2003, 107, 14231–14239.<br />

280 R. Argazzi, C. A. Bignozzi, T. A. Heimer and G. J. Meyer, Inorg.<br />

Chem., 1997, 36, 2–3.<br />

281 F. Liu and G. J. Meyer, Inorg. Chem., 2005, 44, 9305–9313.<br />

282 G. M. Hasselmann and G. J. Meyer, Z. Phys. Chem. (Munich),<br />

1999, 212, 39–44.<br />

283 M. K. Nazeeruddin, R. Humphry-Baker, M. Gra¨tzel and<br />

B. A. Murrer, Chem. Commun., 1998, 719–720.<br />

284 A. Morandeira, I. Lopez-Duarte, M. V. Martinez-Diaz,<br />

B. O’Regan, C. Shuttle, N. A. Haji-Zainulabidin, T. Torres,<br />

E. Palomares and J. R. Durrant, J. Am. Chem. Soc., 2007, 129,<br />

9250–9251.<br />

285 B. C. O’Regan, I. Lopez-Duarte, M. V. Martinez-Diaz,<br />

A. Forneli, J. Albero, A. Morandeira, E. Palomares, T. Torres<br />

and J. R. Durrant, J. Am. Chem. Soc., 2008, 130, 2906–2907.<br />

286 S. Ferrere and B. A. Gregg, J. Am. Chem. Soc., 1998, 120,<br />

843–844.<br />

287 A. Islam, K. Hara, L. P. Singh, R. Katoh, M. Yanagida,<br />

S. Murata, Y. Takahashi, H. Sugihara and H. Arakawa, Chem.<br />

Lett., 2000, 29, 490–491.<br />

288 T. A. Heimer, E. J. Heilweil, C. A. Bignozzi and G. J. Meyer,<br />

J. Phys. Chem. A, 2000, 104, 4256–4262.<br />

289 J.-E. Moser and M. Gra¨tzel, Chimia, 1998, 52, 160–162.<br />

290 F. Liu and G. J. Meyer, Inorg. Chem., 2003, 42, 7351–7353.<br />

291 D. P. Colombo and R. M. Bowman, J. Phys. Chem., 1995, 99,<br />

11752–11756.<br />

292 D. P. Colombo, K. A. Roussel, J. Saeh, D. E. Skinner,<br />

J. J. Cavaleri and R. M. Bowman, Chem. Phys. Lett., 1995,<br />

232, 207–214.<br />

293 G. Rothenberger, J. Moser, M. Gra¨tzel, N. Serpone and<br />

D. K. Sharma, J. Am. Chem. Soc., 1985, 107, 8054–8059.<br />

294 N. Serpone, D. Lawless, R. Khairutdinov and E. Pelizzetti,<br />

J. Phys. Chem., 1995, 99, 16655–16661.<br />

295 D. E. Skinner, D. P. Colombo, J. J. Cavaleri and R. M. Bowman,<br />

J. Phys. Chem., 1995, 99, 7853–7856.<br />

296 W. Shockley and H. J. Queisser, J. Appl. Phys., 1961, 32, 510–519.<br />

297 R. T. Ross and A. J. Nozik, J. Appl. Phys., 1982, 53,<br />

3813–3818.<br />

298 M. Thoss, I. Kondov and H. Wang, Chem. Phys., 2004, 304,<br />

169–181.<br />

299 P. G. Hoertz, D. W. Thompson, L. A. Friedman and G. J. Meyer,<br />

J. Am. Chem. Soc., 2002, 124, 9690–9691.<br />

300 F. Liu, M. Yang and G. J. Meyer, Molecule-to-Particle Charge<br />

Transfer in Sol–Gel Materials, in Handbook of Sol–Gel Science<br />

and Technology: Processing Characterization and Application, ed.<br />

R. M. Almeida, Kluwer Academic Publishers, 2005, vol. II:<br />

Characterization of Sol–Gel Materials and Products,<br />

pp. 400–428.<br />

301 E. Vrachnou, M. Gra¨tzel and A. J. McEvoy, J. Electroanal.<br />

Chem., 1989, 258, 193–205.<br />

302 E. Vrachnou, N. Vlachopoulos and M. Gra¨tzel, J. Chem. Soc.,<br />

Chem. Commun., 1987, 868–870.<br />

303 R. L. Blackbourn, C. S. Johnson and J. T. Hupp, J. Am. Chem.<br />

Soc., 1991, 113, 1060–1062.<br />

304 H. Lu, J. N. Prieskorn and J. T. Hupp, J. Am. Chem. Soc., 1993,<br />

115, 4927–4928.<br />

305 S. F. Fischer and R. P. Van Duyne, Chem. Phys., 1977, 26, 9–16.<br />

306 J. J. Hopfield, Proc. Natl. Acad. Sci. U. S. A., 1974, 71,<br />

3640–3644.<br />

307 J. Jortner, J. Chem. Phys., 1976, 64, 4860–4867.<br />

308 N.R. Kestner, J. Logan and J. Jortner, J. Phys. Chem., 1974, 78,<br />

2148–2166.<br />

309 J. Ulstrup and J. Jortner, J. Chem. Phys., 1975, 63,<br />

4358–4368.<br />

310 R. P. Van Duyne and S. F. Fischer, Chem. Phys., 1974, 5,<br />

183–197.<br />

311 M. Yang, D. W. Thompson and G. J. Meyer, Inorg. Chem., 2000,<br />

39, 3738–3739.<br />

312 A. Zaban, S. Ferrere and B. A. Gregg, J. Phys. Chem. B, 1998,<br />

102, 452–460.<br />

313 A. Zaban, S. Ferrere, J. Sprague and B. A. Gregg, J. Phys. Chem.<br />

B, 1997, 101, 55–57.<br />

314 M. Khoudiakov, A. R. Parise and B. S. Brunschwig, J. Am.<br />

Chem. Soc., 2003, 125, 4637–4642.<br />

315 J. A. Harris, K. Trotter and B. S. Brunschwig, J. Phys. Chem. B,<br />

2007, 111, 6695–6702.<br />

316 S. Verma, P. Kar, A. Das, D. K. Palit and H. N. Ghosh, J. Phys.<br />

Chem. C, 2008, 112, 2918–2926.<br />

317 H. N. Ghosh, J. Phys. Chem. B, 1999, 103, 10382–10387.<br />

318 P. Qu and G. J. Meyer, Dye-Sensitized Electrodes, in Electron<br />

Transfer in Chemistry, ed. V. Balzani, John Wiley & Sons, NY,<br />

2001, vol. IV, ch. 2, part 2, pp. 355–411.<br />

319 I. Ortmans, C. Moucheron and A. Kirsch-De Mesmaeker,<br />

Coord. Chem. Rev., 1998, 168, 233–271.<br />

320 B. V. Bergeron and G. J. Meyer, J. Phys. Chem. B, 2003, 107,<br />

245–254.<br />

321 D. W. Thompson, C. A. Kelly, F. Farzad and G. J. Meyer,<br />

Langmuir, 1999, 15, 650–653.<br />

322 C. J. Kleverlaan, M. T. Indelli, C. A. Bignozzi, L. Pavanin,<br />

F. Scandola, G. M. Hasselman and G. J. Meyer, J. Am. Chem.<br />

Soc., 2000, 122, 2840–2849.<br />

323 B. V. Bergeron, A. Marton, G. Oskam and G. J. Meyer, J. Phys.<br />

Chem. B, 2005, 109, 937–943.<br />

324 P. Wang, B. Wenger, R. Humphry-Baker, J. E. Moser,<br />

J. Teuscher, W. Kantlehner, J. Mezger, E. V. Stoyanov,<br />

S. M. Zakeeruddin and M. Grätzel, J. Am. Chem. Soc., 2005,<br />

127, 6850–6856.<br />

325 D. Gust, T. A. Moore and A. L. Moore, J. Photochem. Photobiol.<br />

B: Biol., 2000, 58, 63–71.<br />

326 T. A. Moore, A. L. Moore and D. Gust, Philos. Trans. R. Soc.<br />

London, B, 2002, 357, 1481–1498.<br />

327 C. S. Christ, J. Yu, X. Zhao, G. T. R. Palmore and<br />

M. S. Wrighton, Inorg. Chem., 1992, 31, 4439–4440.<br />

328 R. Argazzi, C. A. Bignozzi, T. A. Heimer, F. N. Castellano<br />

and G. J. Meyer, J. Am. Chem. Soc., 1995, 117, 11815–11816.<br />

329 R. Argazzi, C. A. Bignozzi, T. A. Heimer, F. N. Castellano and<br />

G. J. Meyer, J. Phys. Chem. B, 1997, 101, 2591–2597.<br />

330 P. Bonhote, J. E. Moser, R. Humphry-Baker, N. Vlachopoulos,<br />

S. M. Zakeeruddin, L. Walder and M. Grätzel, J. Am. Chem.<br />

Soc., 1999, 121, 1324–1336.<br />

331 N. Hirata, J.-J. Lagref, E. J. Palomares, J. R. Durrant,<br />

M. K. Nazeeruddin, M. Gra¨tzel and D. Di Censo, Chem. Eur.<br />

J., 2004, 10, 595–602.<br />

332 S. A. Haque, S. Handa, K. Peter, E. Palomares, M. Thelakkat<br />

and J. R. Durrant, Angew. Chem., Int. Ed., 2005, 44, 5740–5744.<br />

333 H. Wolpher, S. Sinha, J. Pan, A. Johansson, M. J. Lundqvist,<br />

P. Persson, R. Lomoth, J. Bergquist, L. Sun, V. Sundström,<br />

B. Akermark and T. Polivka, Inorg. Chem., 2007, 46, 638–651.<br />

334 C. Kleverlaan, M. Alebbi, R. Argazzi, C. A. Bignozzi,<br />

G. M. Hasselmann and G. J. Meyer, Inorg. Chem., 2000, 39,<br />

1342–1343.<br />

162 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009


335 A. C. Lees, C. J. Kleverlaan, C. A. Bignozzi and J. G. Vos, Inorg.<br />

Chem., 2001, 40, 5343–5349.<br />

336 Y. Xu, G. Eilers, M. Borgstro¨m, J. Pan, M. Abrahamsson,<br />

A. Magnuson, R. Lomoth, J. Bergquist, T. Polivka, L. Sun,<br />

V. Sundström, S. Styring, L. Hammarstro¨m and B. A˚ kermark,<br />

Chem. Eur. J., 2005, 11, 7305–7314.<br />

337 B. Gholamkhass, K. Koike, N. Negishi, H. Hori, T. Sano and<br />

K. Takeuchi, Inorg. Chem., 2003, 42, 2919–2932.<br />

338 D. M. Stanbury, Reduction Potentials Involving Inorganic Free<br />

Radicals in Aqueous Solution, Adv. Inorg. Chem., 1989, 33,<br />

69–138.<br />

339 G. Nord, B. Pedersen and O. Farver, Inorg. Chem., 1978, 17,<br />

2233–2238.<br />

340 G. Nord, B. Pedersen, L. Floryan and P. Pagsberg, Inorg. Chem.,<br />

1982, 21, 2327–2330.<br />

341 G. Nord, Comm. Inorg. Chem., 1992, 13, 221–239.<br />

342 W. K. Wilmarth, D. M. Stanbury, J. E. Byrd, H. N. Po and<br />

C.-P. Chua, Coord. Chem. Rev., 1983, 51, 155–179.<br />

343 B. J. Walter and C. M. Elliott, Inorg. Chem., 2001, 40, 5924–5927.<br />

344 I. V. Nelson and R. T. Iwamoto, J. Electroanal. Chem., 1964, 7,<br />

218–221.<br />

345 J. Desbarres, Bull. Soc. Chim. Fr., 1961, 28, 502.<br />

346 A. J. Parker, J. Chem. Soc. A, 1966, 220–228.<br />

347 R. Alexander, E. C. F. Ko, Y. C. Mac and A. J. Parker, J. Am.<br />

Chem. Soc., 1967, 89, 3703–3712.<br />

348 F. G. K. Baucke, R. Bertram and K. Cruse, J. Electroanal. Chem.,<br />

1971, 32, 247–256.<br />

349 R. Guidelli and G. Piccardi, Electrochim. Acta, 1967, 12, 1085–1095.<br />

350 L. I. Katzin and E. Gebert, J. Am. Chem. Soc., 1955, 77,<br />

5814–5819.<br />

351 A. I. Popov and D. H. Geske, J. Am. Chem. Soc., 1958, 80,<br />

1340–1352.<br />

352 V. A. Macagno, M. C. Giordano and A. J. Arvía, Electrochim.<br />

Acta, 1969, 14, 335–357.<br />

353 X. Wang and D. M. Stanbury, Inorg. Chem., 2006, 45, 3415–3423.<br />

354 J. N. Demas and J. W. Addington, J. Am. Chem. Soc., 1976, 98,<br />

5800–5806.<br />

355 J. N. Demas, J. W. Addington, S. H. Peterson and E. W. Harris,<br />

J. Phys. Chem., 1977, 81, 1039–1043.<br />

356 C. Nasr, S. Hotchandani and P. V. Kamat, J. Phys. Chem. B,<br />

1998, 102, 4944–4951.<br />

357 C. C. Clark, A. Marton and G. J. Meyer, Inorg. Chem., 2005, 44,<br />

3383–3385.<br />

358 J. M. Gardner and G. J. Meyer, J. Am. Chem. Soc., 2008, in press.<br />

359 D. J. Fitzmaurice and H. Frei, Langmuir, 1991, 7, 1129–1137.<br />

360 J. N. Clifford, E. Palomares, M. K. Nazeeruddin, M. Gra¨tzel and<br />

J. R. Durrant, J. Phys. Chem. C, 2007, 111, 6561–6567.<br />

361 S. Pelet, J. E. Moser and M. Gra¨tzel, J. Phys. Chem. B, 2000, 104,<br />

1791–1795.<br />

362 P. J. Cameron, L. M. Peter, S. M. Zakeeruddin and M. Gra¨tzel,<br />

Coord. Chem. Rev., 2004, 248, 1447–1453.<br />

363 B. A. Gregg, F. Pichot, S. Ferrere and C. L. Fields, J. Phys.<br />

Chem. B, 2001, 105, 1422–1429.<br />

364 J. Desilvestro, M. Grätzel, L. Kavan, J. Moser and<br />

J. Augustynski, J. Am. Chem. Soc., 1985, 107, 2988–2990.<br />

365 N. Vlachopoulos, P. Liska, J. Augustynski and M. Grätzel,<br />

J. Am. Chem. Soc., 1988, 110, 1216–1220.<br />

366 G. Oskam, B. V. Bergeron, G. J. Meyer and P. C. Searson,<br />

J. Phys. Chem. B, 2001, 105, 6867–6873.<br />

367 P. Wang, S. M. Zakeeruddin, J. E. Moser, R. Humphry-Baker<br />

and M. Gra¨tzel, J. Am. Chem. Soc., 2004, 126, 7164–7165.<br />

368 Z. Zhang, P. Chen, T. N. Murakami, S. M. Zakeeruddin and<br />

M. Grätzel, Adv. Funct. Mater., 2008, 18, 341–346.<br />

369 Z. Zhang, P. C. Chen, S. M. Zakeeruddin, J.-E. Moser and<br />

M. Grätzel, Mater. Res. Soc. Symp. Proc., 2007, 1013E, Z06-02.<br />

370 N. Sutin and C. Creutz, J. Chem. Educ., 1983, 60, 809–814.<br />

371 H. Nusbaumer, J. E. Moser, S. M. Zakeeruddin,<br />

M. K. Nazeeruddin and M. Grätzel, J. Phys. Chem. B, 2001,<br />

105, 10461–10464.<br />

372 H. Nusbaumer, S. M. Zakeeruddin, J.-E. Moser and M. Gra¨tzel,<br />

Chem. Eur. J., 2003, 9, 3756–3763.<br />

373 S. Nakade, T. Kanzaki, W. Kubo, T. Kitamura, Y. Wada and<br />

S. Yanagida, J. Phys. Chem. B, 2005, 109, 3480–3487.<br />

374 S. Nakade, Y. Makimoto, W. Kubo, T. Kitamura, Y. Wada<br />

and S. Yanagida, J. Phys. Chem. B, 2005, 109, 3488–3493.<br />

375 S. A. Sapp, C. M. Elliott, C. Contado, S. Caramori and<br />

C. A. Bignozzi, J. Am. Chem. Soc., 2002, 124, 11215–11222.<br />

376 M. J. Scott, J. J. Nelson, S. Caramori, C. A. Bignozzi and<br />

C. M. Elliott, Inorg.Chem., 2007, 46, 10071–10078.<br />

377 E. A. Medlycott, I. Theobald and G. S. Hanan, Eur. J. Inorg.<br />

Chem., 2005, 1223–1226.<br />

378 L. X. Chen, G. Jennings, T. Liu, D. J. Gosztola, J. P. Hessler,<br />

D. V. Scaltrito and G. J. Meyer, J. Am. Chem. Soc., 2002, 124,<br />

10861–10867.<br />

379 M. Brugnati, S. Caramori, S. Cazzanti, L. Marchini, R. Argazzi<br />

and C. A. Bignozzi, Int. J. Photoenergy, 2007, 2007, 80756–80765.<br />

380 S. Hattori, Y. Wada, S. Yanagida and S. Fukuzumi, J. Am.<br />

Chem. Soc., 2005, 127, 9648–9654.<br />

381 A. Staniszewski, S. Ardo, Y. Sun, F. N. Castellano and<br />

G. J. Meyer, J. Am. Chem. Soc., 2008, 130, 11586–11587.<br />

382 I. Montanari, J. Nelson and J. R. Durrant, J. Phys. Chem. B,<br />

2002, 106, 12203–12210.<br />

383 R. Kohlrausch, Ann. Phys, Chem. (Poggendorff), 1854, 167(1),<br />

91, 56–82.<br />

384 R. Kohlrausch, Ann. Phys. Chem. (Poggendorff), 1854, 167(2),<br />

91, 179–214.<br />

385 G. Williams and D. C. Watts, Trans. Faraday Soc., 1970, 66, 80–85.<br />

386 T. A. Heimer and G. J. Mayer, J. Lumin., 1996, 70, 468–478.<br />

387 A. Furube, R. Katoh, K. Hara, T. Sato, S. Murata, H. Arakawa<br />

and M. Tachiya, J. Phys. Chem. B, 2005, 109, 16406–16414.<br />

388 M. Gra¨tzel, Cattech, 1999, 3, 4–17.<br />

389 G. M. Hasselmann and G. J. Meyer, J. Phys. Chem. B, 1999, 103,<br />

7671–7675.<br />

390 D. Kuciauskas, M. S. Freund, H. B. Gray, J. R. Winkler and<br />

N. S. Lewis, J. Phys. Chem. B, 2001, 105, 392–403.<br />

391 E. A. Schiff, Phys. Rev. B, 1981, 24, 6189.<br />

392 F. W. Schmidlin, Philos. Mag. B, 1980, 41, 535–570.<br />

393 T. Tiedje and A. Rose, Solid State Commun., 1981, 37,<br />

49–52.<br />

394 G. Pfister and H. Scher, Phys. Rev. B, 1977, 15, 2062.<br />

395 H. Scher and E. W. Montroll, Phys. Rev. B, 1975, 12, 2455.<br />

396 R. Argazzi, C. A. Bignozzi, T. A. Heimer, F. N. Castellano and<br />

G. J. Meyer, Inorg. Chem., 1994, 33, 5741–5749.<br />

397 C. P. Lindsey and G. D. Patterson, J. Chem. Phys., 1980, 73,<br />

3348–3357.<br />

398 D. L. Huber, Phys. Rev. E, 1996, 53, 6544–6546.<br />

399 M. O. Vlad and M. C. Mackey, J. Math. Phys., 1995, 36, 1834–1853.<br />

400 A. V. Barzykin and M. Tachiya, J. Phys. Chem. B, 2002, 106,<br />

4356–4363.<br />

401 R. Katoh, A. Furube, A. V. Barzykin, H. Arakawa and<br />

M. Tachiya, Coord. Chem. Rev., 2004, 248, 1195–1213.<br />

402 Stretched exponential function, http://en.wikipedia.org/wiki/<br />

Kohlrausch-Williams-Watts_function.<br />

403 J. Bisquert, J. Phys. Chem. C, 2007, 111, 17163–17168.<br />

404 R. G. Palmer, D. L. Stein, E. Abrahams and P. W. Anderson,<br />

Phys. Rev. Lett., 1984, 53, 958–961.<br />

405 M. F. Shlesinger, Phys. Today, 1991, 44, 15–94.<br />

406 A. Plonka, J. Mol. Liq., 1995, 64, 39–48.<br />

407 F. N. Castellano, T. A. Heimer, M. T. Tandhasetti and<br />

G. J. Meyer, Chem. Mater., 1994, 6, 1041–1048.<br />

408 L. Forro, O. Chauvet, D. Emin, L. Zuppiroli, H. Berger and<br />

F. Levy, J. Appl. Phys., 1994, 75, 633–635.<br />

409 E. Hendry, F. Wang, J. Shan, T. F. Heinz and M. Bonn, Phys.<br />

Rev. B, 2004, 69, 081101.<br />

410 N. Kopidakis, E. A. Schiff, N. G. Park, J. van de Lagemaat and<br />

A. J. Frank, J. Phys. Chem. B, 2000, 104, 3930–3936.<br />

411 E. Hendry, M. Koeberg, B. O’Regan and M. Bonn, Nano Lett.,<br />

2006, 6, 755–759.<br />

412 R. Könenkamp, Phys. Rev. B, 2000, 61, 11057.<br />

413 L. M. Peter, J. Phys. Chem. C, 2007, 111, 6601–6612.<br />

414 J. A. Anta, I. Mora-Sero, T. Dittrich and J. Bisquert, Phys. Chem.<br />

Chem. Phys., 2008, 10, 4478–4485.<br />

415 J. Bisquert, J. Phys. Chem. B, 2004, 108, 2323–2332.<br />

416 J. Bisquert and V. S. Vikhrenko, J. Phys. Chem. B, 2004, 108,<br />

2313–2322.<br />

417 L. A. Lyon and J. T. Hupp, J. Phys. Chem., 1995, 99,<br />

15718–15720.<br />

418 S. Sodergren, H. Siegbahn, H. Rensmo, H. Lindstrom,<br />

A. Hagfeldt and S. E. Lindquist, J. Phys. Chem. B, 1997, 101,<br />

3087–3090.<br />

This journal is c The Royal Society of Chemistry 2009 Chem.Soc.Rev., 2009, 38, 115–164 | 163


419 M. Wagemaker, A. P. M. Kentgens and F. M. Mulder, Nature,<br />

2002, 418, 397–399.<br />

420 M. Wagemaker, R. van de Krol, A. P. M. Kentgens, A. A. van<br />

Well and F. M. Mulder, J. Am. Chem. Soc., 2001, 123,<br />

11454–11461.<br />

421 D. Nister, K. Keis, S.-E. Lindquist and A. Hagfeldt, Sol. Energy<br />

Mater., 2002, 73, 411–423.<br />

422 J. L. Scales and A. L. Ward, J. Appl. Phys., 1968, 39, 1692–1700.<br />

423 W. van Roosbroeck, Phys. Rev., 1953, 91, 282–289.<br />

424 F. Cao, G. Oskam, G. Meyer and P. Searson, J. Phys. Chem.,<br />

1996, 100, 17021–17027.<br />

425 A. Solbrand, H. Lindstrom, H. Rensmo, A. Hagfeldt,<br />

S. E. Lindquist and S. Sodergren, J. Phys. Chem. B, 1997, 101,<br />

2514–2518.<br />

426 B. B. Smith and C. A. Koval, J. Electroanal. Chem., 1990, 277,<br />

43–72.<br />

427 A. Solbrand, A. Henningsson, S. Sodergren, H. Lindstrom,<br />

A. Hagfeldt and S. E. Lindquist, J. Phys. Chem. B, 1999, 103,<br />

1078–1083.<br />

428 J. van de Lagemaat and A. J. Frank, J. Phys. Chem. B, 2001, 105,<br />

11194–11205.<br />

429 S. Nakade, S. Kambe, T. Kitamura, Y. Wada and S. Yanagida,<br />

J. Phys. Chem. B, 2001, 105, 9150–9152.<br />

430 S. Kambe, S. Nakade, T. Kitamura, Y. Wada and S. Yanagida,<br />

J. Phys. Chem. B, 2002, 106, 2967–2972.<br />

431 H. G. Agrell, G. Boschloo and A. Hagfeldt, J. Phys. Chem. B,<br />

2004, 108, 12388–12396.<br />

432 B. C. O’Regan, K. Bakker, J. Kroeze, H. Smit, P. Sommeling and<br />

J. R. Durrant, J. Phys. Chem. B, 2006, 110, 17155–17160.<br />

433 B. C. O’Regan and J. R. Durrant, J. Phys. Chem. B, 2006, 110,<br />

8544–8547.<br />

434 G. Boschloo and A. Hagfeldt, J. Phys. Chem. B, 2005, 109,<br />

12093–12098.<br />

435 A. Goossens, B. van der Zanden and J. Schoonman, Chem. Phys.<br />

Lett., 2000, 331, 1–6.<br />

436 N. Kopidakis, K. D. Benkstein, J. van de Lagemaat, A. J. Frank,<br />

Q. Yuan and E. A. Schiff, Phys. Rev. B, 2006, 73, 045326.<br />

437 L. M. Peter, A. B. Walker, G. Boschloo and A. Hagfeldt, J. Phys.<br />

Chem. B, 2006, 110, 13694–13699.<br />

438 J. E. Moser and M. Grätzel, Chem. Phys., 1993, 176, 493–500.<br />

439 P. Liska, N. Vlachopoulos, M. K. Nazeeruddin, P. Comte and<br />

M. Grätzel, J. Am. Chem. Soc., 1988, 110, 3686–3687.<br />

440 Q. Wang, S. M. Zakeeruddin, M. K. Nakeeruddin, R. Humphry-<br />

Baker and M. Gra¨tzel, J. Am. Chem. Soc., 2006, 128, 4446–4452.<br />

441 M. Chou, C. Creutz and N. Sutin, J. Am. Chem. Soc., 1977, 99,<br />

5615–5623.<br />

442 R. C. Young, F. R. Keene and T. J. Meyer, J. Am. Chem. Soc.,<br />

1977, 99, 2468–2473.<br />

443 A. G. Motten, K. Hanck and M. K. DeArmound, Chem. Phys.<br />

Lett., 1981, 79, 541–546.<br />

444 G. A. Heath, L. J. Yellowlees and P. S. Braterman, Chem. Phys.<br />

Lett., 1982, 92, 646–648.<br />

445 A. Staniszewski, A. J. Morris, T. Ito and G. J. Meyer, J. Phys.<br />

Chem. B, 2007, 111, 6822–6828.<br />

446 R. A. Marcus, J. Chem. Phys., 1956, 24, 966–978.<br />

447 R. A. Marcus, Discuss. Faraday Soc., 1960, 29, 21–31.<br />

448 A. B. P. Lever, Ligand Electrochemical Parameters and<br />

Electrochemical-optical Relationships, in Comprehensive Coordination<br />

Chemistry, II, Elsevier Science, 2003, vol. 2, pp. 251–268.<br />

449 X. Dang and J. T. Hupp, J. Am. Chem. Soc., 1999, 121, 8399–8400.<br />

450 J. N. Clifford, E. Palomares, M. K. Nazeeruddin, M. Grätzel,<br />

J. Nelson, X. Li, N. J. Long and J. R. Durrant, J. Am. Chem.<br />

Soc., 2004, 126, 5225–5233.<br />

451 S. G. Yan and J. T. Hupp, J. Phys. Chem., 1996, 100, 6867–6870.<br />

452 S. G. Yan and J. T. Hupp, J. Phys. Chem. B, 1997, 101,<br />

1493–1495.<br />

453 S. G. Yan, J. S. Prieskorn, Y. Kim and J. T. Hupp, J. Phys. Chem.<br />

B, 2000, 104, 10871–10877.<br />

454 P. M. Sommeling, B. C. O’Regan, R. R. Haswell, H. J. P. Smit,<br />

N. J. Bakker, J. J. T. Smits, J. M. Kroon and J. A. M. van<br />

Roosmalen, J. Phys. Chem. B, 2006, 110, 19191–19197.<br />

455 D. A. Gaal and J. T. Hupp, J. Am. Chem. Soc., 2000, 122,<br />

10956–10963.<br />

456 S. A. Haque, Y. Tachibana, R. L. Willis, J. E. Moser, M. Grätzel,<br />

D. R. Klug and J. R. Durrant, J. Phys. Chem. B, 2000, 104,<br />

538–547.<br />

457 S. A. Haque, Y. Tachibana, D. R. Klug and J. R. Durrant,<br />

J. Phys. Chem. B, 1998, 102, 1745–1749.<br />

458 D. Wang, R. Mendelsohn, E. Galoppini, P. G. Hoertz,<br />

R. A. Carlisle and G. J. Meyer, J. Phys. Chem. B, 2004, 108,<br />

16642–16653.<br />

459 Saif A. Haque, Jong S. Park, M. Srinivasarao and J. R. Durrant,<br />

Adv. Mater., 2004, 16, 1177–1181.<br />

460 S. Handa, H. Wietasch, M. Thelakkat, J. R. Durrant and<br />

S. A. Haque, Chem. Commun., 2007, 1725–1727.<br />

461 J. N. Clifford, G. Yahioglu, L. R. Milgrom and J. R. Durrant,<br />

Chem. Commun., 2002, 1260–1261.<br />

462 A. V. Barzykin and M. Tachiya, J. Phys. Chem. B, 2004, 108,<br />

8385–8389.<br />

463 A. N. M. Green, R. E. Chandler, S. A. Haque, J. Nelson and<br />

J. R. Durrant, J. Phys. Chem. B, 2005, 109, 142–150.<br />

464 S. Y. Huang, G. Schlichthorl, A. J. Nozik, M. Gra¨tzel and<br />

A. J. Frank, J. Phys. Chem. B, 1997, 101, 2576–2582.<br />

465 S. Sodergren, A. Hagfeldt, J. Olsson and S.-E. Lindquist, J. Phys.<br />

Chem., 1994, 98, 5552–5556.<br />

466 M. L. Rosenbluth and N. S. Lewis, J. Phys. Chem., 1989, 93,<br />

3735–3740.<br />

467 A. Kumar, P. G. Santangelo and N. S. Lewis, J. Phys. Chem.,<br />

1992, 96, 834–842.<br />

468 M. Gra¨tzel and K. Kalyanasundaram, Curr. Sci., 1994, 66,<br />

706–714.<br />

469 S. Altobello, R. Argazzi, S. Caramori, C. Contado, S. D. Fre,<br />

P. Rubino, C. Chone, G. Larramona and C. A. Bignozzi, J. Am.<br />

Chem. Soc., 2005, 127, 15342–15343.<br />

470 M. Yanagida, T. Yamaguchi, M. Kurashige, K. Hara, R. Katoh,<br />

H. Sugihara and H. Arakawa, Inorg. Chem., 2003, 42, 7921–7931.<br />

471 C. C. Clark, G. J. Meyer, Q. Wei and E. Galoppini, J. Phys.<br />

Chem. B, 2006, 110, 11044–11046.<br />

472 C. C. Clark, A. Marton, R. Srinivasan, A. A. Narducci Sarjeant<br />

and G. J. Meyer, Inorg. Chem., 2006, 45, 4728–4734.<br />

473 L. M. Peter and K. G. U. Wijayantha, Electrochim. Acta, 2000,<br />

45, 4543–4551.<br />

474 N. W. Duffy, L. M. Peter, R. M. G. Rajapakse and K. G. U.<br />

Wijayantha, J. Phys. Chem. B, 2000, 104, 8916–8919.<br />

475 N. W. Duffy, L. M. Peter, R. M. G. Rajapakse and K. G. U.<br />

Wijayantha, Electrochem. Commun., 2000, 2, 658–662.<br />

476 L. M. Peter, N. W. Duffy, R. L. Wang and K. G. U. Wijayantha,<br />

J. Electroanal. Chem., 2002, 524–525, 127–136.<br />

477 N. W. Duffy, L. M. Peter and K. G. U. Wijayantha, Electrochem.<br />

Commun., 2000, 2, 262–266.<br />

478 G. Schlichthorl, S. Y. Huang, J. Sprague and A. J. Frank, J. Phys.<br />

Chem. B, 1997, 101, 8141–8155.<br />

479 G. Schlichthorl, N. G. Park and A. J. Frank, J. Phys. Chem. B,<br />

1999, 103, 782–791.<br />

480 N. Kopidakis, K. D. Benkstein, J. vandeLagemaat and<br />

A. J. Frank, J. Phys. Chem. B, 2003, 107, 11307–11315.<br />

481 Y. Liu, A. Hagfeldt, X.-R. Xiao and S.-E. Lindquist, Sol. Energy<br />

Mater., 1998, 55, 267–281.<br />

482 C. Bauer, G. Boschloo, E. Mukhtar and A. Hagfeldt, J. Phys.<br />

Chem. B, 2002, 106, 12693–12704.<br />

483 H. van’t Spijker, B. O’Regan and A. Goossens, J. Phys. Chem. B,<br />

2001, 105, 7220–7226.<br />

164 | Chem.Soc.Rev., 2009, 38, 115–164 This journal is c The Royal Society of Chemistry 2009

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!