Cronfa - Swansea University Open Access Repository
_____________________________________________________________
This is an author produced version of a paper published in:
Science Progress
Cronfa URL for this paper:
http://cronfa.swan.ac.uk/Record/cronfa34123
_____________________________________________________________
Paper:
De Castro, C., Dimitrov, S., Burrows, H., Douglas, P. & Davies, M. (2017). Photoinduced charge transfer: from
photography to solar energy. Science Progress, 100(2), 212-230.
http://dx.doi.org/10.3184/003685017X14901006155099
_____________________________________________________________
This item is brought to you by Swansea University. Any person downloading material is agreeing to abide by the terms
of the repository licence. Copies of full text items may be used or reproduced in any format or medium, without prior
permission for personal research or study, educational or non-commercial purposes only. The copyright for any work
remains with the original author unless otherwise specified. The full-text must not be sold in any format or medium
without the formal permission of the copyright holder.
Permission for multiple reproductions should be obtained from the original author.
Authors are personally responsible for adhering to copyright and publisher restrictions when uploading content to the
repository.
http://www.swansea.ac.uk/iss/researchsupport/cronfa-support/
Photoinduced Charge Transfer: from Photography
to Solar Energy
CATHERINE S. DE CASTRO, STOICHKO DIMITROV, HUGH D. BURROWS, PETER
DOUGLAS, and MATTHEW L. DAVIES
Catherine S. De Castro received her PhD with specialisation in
Photochemistry in 2016 from the University of Coimbra, Portugal. She is
currently working as a Post-Doctoral Research Assistant at SPECIFIC in
Swansea University. Her research interests include photophysical
characterisation of photovoltaic related materials, chemosensors and probes.
E-mail: C.S.DeCastro@swansea.ac.uk
Stoichko Dimitrov is a Sêr Cymru II fellow at the College of Engineering of
Swansea University and SPECIFIC. His research interests are in the excitedstates of organic semiconductors and water photocatalysis. His PhD research
(Boston College, USA, 2010) focused on femtosecond spectroscopy of
semiconducting nanocrystals and DNA.
E-mail: Stoichko.Dimitrov@Swansea.ac.uk
Hugh Burrows is Emeritus Professor of Chemistry, University of Coimbra,
Portugal. He is a native of England and did his B.Sc. (London) and Ph.D.
(Sussex) there. He has been in Portugal for over 30 years, and his research
has concentrated on various aspects of materials chemistry, polymers and
photochemistry.
E-mail: burrows@ci.uc.pt
Peter Douglas is Associate Professor in Chemistry in the College of
Medicine, Swansea University and Honorary Professor at the University of
KwaZulu-Natal. His research interests are in applied physical chemistry,
particularly photochemistry.
E-mail: P.Douglas@swansea.ac.uk
Dr Matthew L. Davies is a Senior Lecturer and head of the Applied
Photochemistry Group in the Materials Research Centre, College of
Engineering at Swansea University. His research focuses on the
photochemistry of materials for low-cost photovoltaic applications, with the
ultimate aim of improving stability and performance.
E-mail: m.l.davies@swansea.ac.uk
ABSTRACT
To celebrate the centenary of Science Progress we offer a short survey of the progress made
over the past one hundred years in the research and application of photoinduced charge transfer.
After a brief historical overview and introduction to photoinduced charge transfer, we discuss
developments in the theory and practice of photography, photovoltaics, photocatalysis,
fluorescent probes and chemosensing.
Keywords: chemosensing, fluorescent probes, photocatalysis, photography, photoinduced
charge transfer, photovoltaics, solar energy.
1. Introduction
1.1.Historical overview
In 1917 the world of atomic chemistry was that of the Old Quantum Theory of Planck, Einstein,
Bohr and Sommerfeld. The electron had been known for twenty years, identified as the particle
of cathode rays by Thompson in 1897. Planck had introduced quantisation in his theory of the
emission spectrum of a ‘black-body’ radiator in 1900; Einstein used ‘light quanta’ to explain
the photoelectric effect and the low temperature heat capacity of solids; and Bohr, in 1913
‘solved’ the problem of atomic structure thrown up by Rutherford’s studies of alpha particle
scattering by combining Rutherford’s model with Planck’s quantisation.1
Within 10 years, the Old Quantum Theory was replaced by the new quantum (or wave)
mechanics of Schrödinger and Heisenberg, and the structure of atoms and molecules and the
interaction between matter and light had been described in much the same terms we use
today.2,3
But the experimental study of the interaction between light and chemical change and light and
electricity already had a long history in 1917. Two hundred years earlier, Schulze had observed
the light sensitivity of silver salts and used it to make images of stencilled words on chalk
containing a little silver nitrate, although he was unable to make the image permanent. The
chemical effects of visible and ultraviolet light were studied by numerous scientist in the
following hundred years.4 The first permanent photographic image, which is not silver halide
based but rather a bitumen photoresist, is due to Niépce in 1826, and Daguerre and Fox-Talbot
introduced their Daguerreotype and Calotype silver halide photographic processes in 1839 and
1841, respectively. At around the same time, 1839, Becquerel created the world's first
photovoltaic (PV) cell from an illuminated silver chloride electrode in an acidic solution.5 By
1917, technologies which we know now involve photoinduced charge separation were in
operation across the world with every click of a camera shutter and in every engineer, architect,
and pattern-maker’s blue-print. Figure 1 give a time-line of significant developments related to
photoinduced charge transfer.6
2
Figure 1. Time-line of developments related to photoinduced charge transfer.
2. Concepts and Applications
We now interpret atomic and molecular structure in terms of the electronic orbitals of the
Schrodinger equation. The absorption of a photon (light ‘particle’ of energy hv where h is
Planck’s constant and v the frequency) causes the transition of an electron from one orbital to
a higher energy orbital to give an excited-state. The excited-state has a different electronic
distribution to that of the ground-state. In some cases, this can involve quite significant charge
separation, and, since the excited-state also has considerable excess energy over the groundstate, its chemistry is often very different from that of the ground-state. Our discussion here
will concentrate on processes involving the photoinduced movement of charge, charge transfer.
Charge transfer (CT) is sometimes used in a broad sense referring to the transfer of: energy,
charge, electrons and/or ions,7 it can be intra or intermolecular. Excited-states have lower
ionisation potentials and greater electron affinities than their corresponding ground-states.
Thus the probability that they react by charge transfer processes is greater than that for groundstate molecules. Table 1 shows the general processes involved, while Figure 2Figure 1 gives
pictorial representations of the key processes of energy and electron transfer.
3
Table 1. Photoexcitation and subsequent charge transfer processes between A and B, which may or
may not be linked together by L.
Processes
Photoexcitation
Charge transfer
Oxidative
Reductive
Electron
transfer
Energy transfer
Intermolecular
A + hv A*
B + hv B*
A* + B ↔ (A...B)*
B* + A ↔ (B…A)*
A* + B A+ + BA* + B A- + B+
A* + B A + B*
Intramolecular
A-L-B + hv A*-L-B
A-L-B + hv A-L-B*
A*-L-B A+-L-BA*-L-B A+-L + BA*-L- B A--L-B+
A*-L- B A--L + B+
A*-L-B A-L-B*
Figure 2. Pictorial representations of radiative, coulombic/Förster and exchange energy transfer,
and electron transfer, adapted from refs. 8,9. HOMO and LUMO are the highest occupied and the
lowest unoccupied molecular orbitals, respectively. Note that Figure 2 is a simplified schematic; we
have not assigned spins to the electrons, although this is a topic we discuss towards the end of this
review.
4
2.1.Photography
Fox-Talbot’s 1841 Calotype process,10 laid the basis of the positive-negative photographic
process involving the ‘capturing’ of a latent image in a silver halide ‘emulsion’ of
microcrystals, followed by chemical ‘development’ and ‘fixing’ to give an image. By 1917
many improvements had been made to give more sensitive emulsions more stable prints, and
easier to use processes.11,12 The most important technological advance for the discussion here
was dye sensitisation.
Silver halides absorb in the UV and blue spectral region but not much beyond, so early
photographs lack a green and red response, and therefore sometimes give very obvious poor
tonal rendition. In fact, Maxwell’s famous 1861 lecture at the Royal Institution where he
demonstrated his theory of ‘three colour reproduction’ by superposition of projections through
three separate images of a brightly coloured tartan ribbon recorded in the red, green, and blue
spectral region made in collaboration with the photographer Sutton13 only just about worked
because the filter dyes used also give different UV transmissions and fabric dyes absorb
differently in the UV as well as the visible.14 Dye sensitisation,15 (the primary mechanism of
which we now know is electron injection into the silver halide crystal from a photoexcited
adsorbed dye), was discovered by Vogel in 1873.16 ‘Dry photographic plates’ using erythrosine
(Figure 5G) as the green sensitising dye were introduced in 1884 and by 1906 ‘panchromatic
plates’ were available which had both red and green sensitising dyes, and which could give a
tonal response across the whole visible spectrum comparable to that experienced by the human
eye.17
So practical photography was in an advance stage by 1917:11 black and white photography was
widespread as an amateur hobby and recording medium for family events (the Kodak Brownie
camera with roll film was introduced in 1900 at a price of $1.00). It was a well-established art
form, and recording medium for news, social commentary, historical record, and exploration.18
Practical colour photography was available through the Lumière Autochrome process, based
on a light filtration ‘screen’ process using a mosaic of red, green, blue, dyed starch grains,
overlain on a silver halide emulsion i.e. essentially black and white made colour through a
matrix of colour micro-filters. Colour motion pictures using additive red, green, blue processes
were being developed, and Lippman had been awarded the 1908 Nobel Prize in Physics "for
his method of reproducing colours photographically based on the phenomenon of
interference".19
However, in photography, like many technologies with immediate commercial application,
scientific understanding lagged behind practical development. Insight into the science of silver
halide photography had to wait for an understanding of the structure and photoelectronic
properties of dyes and semiconductors, and the Gurney-Mott theory of latent image formation
in 1938.20
Silver halides are semiconductors, and the electronic structure and photoelectronic properties
of semiconductors is described in a different way from that in molecules.21 In molecules,
electronic orbitals are localised on molecular species; in semiconductors there are two ‘bands’
of orbitals, which can be considered to be the results of overlap of many molecular orbitals,
which extend across the whole structure: the valence band holds the valence electrons, while
above this, and separated by the semiconductor band-gap, is the conduction band. Direct
absorption causes the transition of an electron from the valence to the conduction band, both
the resultant positively charged hole in the valence band, and the negatively charged electron
5
in the conduction band are mobile and can be trapped at various trap sites in the semiconductor
structure.
Figure 3 shows the general photo-electronic properties of a semiconductor, and the process of
dye sensitisation. In the unsensitised photographic process, the conduction band electron
produced by direct excitation is trapped at a shallow trap site, and subsequently leads to
reduction of an Ag+ to Ag0. Repetition of this process by successive photon absorption events
leads to a speck of at least 4 silver atoms which can catalyse reduction of the whole silver
halide grain by the chemical developer used in the development process. The distribution of
silver grains with and without these catalytic Ag>40 sites is the invisible latent image. In
practise, because of loss processes, it takes at least ~10-15 photons to create one Ag>40 site.22
Figure 3. Direct absorption and dye sensitisation in semiconductors.
In dye sensitisation, the dye, which is present at a very low concentration and is adsorbed onto
the surface of the silver halide grain, absorbs the photon to generate an excited-state which then
injects an electron into the silver halide conduction band. This leaves an oxidised dye molecule,
and while the fate of the oxidised dye is not so important in photography, highly efficient
regeneration of the dye from oxidised dye is essential in dye sensitised solar energy conversion
discussed later.
The ability to make three different emulsions sensitive to blue, green, and red light, also
allowed the possibility of colour reproduction if the colour response could be coupled with a
coloured output, for example in the form of a dye associated with each blue, green, red sensitive
emulsion layer. Great advances in colour photography were made from the 1920’s onwards
until by the mid to latter part of the 1900’s colour photography was a very mature industry
supported by a well understood science.23,24 Much, but not all, colour photography is now made
6
using digital technologies, even so, the ideas developed, in part at least in response to silver
halide photography, of semiconductor structure, photochemistry, and dye sensitised charge
injection, remain key concepts in two of the emerging technologies discussed here, namely,
photo-purification of water by semiconductors, and PV solar energy conversion.
2.2.The ferrioxalate chemical actinometer and blueprints
A chemical actinometer measures the number of photons in a light beam by the chemical
reaction the beam produces. The most useful is probably the ferrioxalate actinometer.25,26
Photolysis of ferrioxalate, [Fe(III)(ox)33-], in solution generates ferrous ions in a high quantum
yield27 by a ligand to metal charge transfer (LMCT) transition. The amount of Fe(II) produced
can be measured spectrophotometrically using phenanthroline (phen) which gives an intensely
coloured red Fe(II)(phen)32+ complex.
The formation of a blueprint involves similar reactions, although the first iron complex used
by Herschel, who introduced the process as the Cyanotype in 1842,10,28 was ferric citrate rather
than ferric oxalate. In the blueprint/Cyanotype process, the paper is sensitised with a mixture
of ferric citrate and potassium ferricyanide and upon exposure to light the photoproduced Fe(II)
ions complex with ferricyanide to give the insoluble pigment Prussian Blue with its
characteristic blueprint colour (Figure 4Error! Reference source not found. shows a modern
Cyanotype).
Figure 4. A modern Cyanotype print of the Paço das Escolas (Scholars' Square) of Coimbra
University. Printed from a negative on photoinkjet paper made from a scanned B&W 35 mm negative.
7
A
2+
B
D
N
3-
N
C
N
N
C
C
III
N
C
N
N
Ru
Fe
C
N
N
C
N
N
C
N
N
N
N
N
Cu
N
N
N
N
E
F
G
OCH3
O
I
NaO
I
I
O
O
I
O
ONa
Figure 5. Structure of (A) ferrycyanide, (B) Ru(bipy)32+, (C) copper phthalocyanine (D) perylene, (E)
fullerene C60 (buckminsterfullerene), (F) PCBM (phenyl-C61-butyric acid methyl ester) and (G)
erythrosine.
2.3.Molecular photoredox reactions, solar energy conversion by water splitting?
The energy crisis of the early 1970’s brought urgency to the development of methods for solar
energy conversion and storage. The scientific community response since then has been an
enormous body of work on ways of converting light energy into thermal, electrical or chemical
energy.
One approach is to catch the photon energy in formation of molecular excited-states and use
excited-state photoredox reactions to split water into H2 and O2. The H2 produced could then
be stored as an environmentally friendly chemical fuel for a ‘hydrogen economy’ to replace
the ‘oil economy’.29,30
Much early work in the field used porphyrins,31 and ruthenium polypyridyl complexes, such as
Ru(bipy)32+ (Figure 5B).32 Ru(bipy)32+ is an orange complex with a relatively long lived charge
transfer triplet state made in ~100% yield, and has ground and excited-state redox couples
which straddle those for water reduction and oxidation. Figure 6 shows the redox energetics
and processes involved in this approach.
While water oxidation and water reduction can be made quite efficient individually using
sacrificial electron donors or acceptors, coupling the two together has, to date, been
unsuccessful in any practical sense. However, the enormous body of research on photoredox
8
reactions of metal complexes since the 1970s has provided a firm foundation for much
subsequent work on photoredox and photocatalytic chemistry, as well as colloidal metal, and
metal oxide, redox-catalysis. In particular, the work on photoredox reactions of Ru(II)
compounds led directly to the development of dye sensitised solar cells.
Figure 6. Energetics of redox processes in water splitting via electron transfer quenching of
*TRu(bipy)32+, where *T signifies the triplet excited state.
2.4. Photoredox reactions at electrodes – solar energy conversion by semiconductor
water splitting, and photocatalytic water and air purification
Another way of photocatalytic water splitting involves light excitation of a semiconducting
material like TiO2 to generate electrons (e-) and holes (h+), followed by a series of charge
transfer steps at the semiconductor-water interface to dissociate water into H2 and O2. In 1970,
Fujishima and Honda reported the first electrochemical cell which successfully evolved H2
from water solutions using TiO2.33 The cell consisted of two compartments (Figure 7) which
housed a TiO2 electrode (where water was oxidised to O2) and a platinum (Pt) counter-electrode
(where water was reduced to H2). Since then much research has been conducted to develop
materials and cells with better solar to hydrogen conversion efficiencies. The key challenges
lie in the difficulty of designing a photocatalyst which can (i) absorb most of the visible solar
spectrum, (ii) dissociate water without the need of much additional applied electrical bias and
(iii) be water stable.
9
Figure 7. Electrochemical cell for water splitting, in which TiO2 is the light absorber and active
water oxidising catalyst. The charges are then used up at the two electrodes for the half-cell reactions
for water splitting.33
In addition to water splitting, photocatalysis has potential in other applications including waste
water and air purification, self-cleaning surfaces and CO2 reduction. For example, for water
treatment, TiO2 is the most frequent choice as a catalyst, because of its ability to drive single
step oxidation of water to hydroxyl radicals.34 Once created, the hydroxyl radicals are
extremely reactive and break most bonds in organic molecules. Whilst waste water treatment
via photocatalysis is not yet commercially viable due to low rates of destruction of molecules,
self-cleaning surfaces employing this process are used in different parts of the world.35
2.5.Photovoltaics
2.5.1. Dye Sensitised Solar Cells
The working mechanism of dye-sensitised solar cells (DSSCs) is analogous to the mechanism
described for the sensitisation of silver halide in photography. In DSSCs charge generation and
transport are separated and facilitated by a dye sensitiser adsorbed on a semiconductor (Figure
3) of an appropriate band gap (usually TiO2, but others such as ZnO have also been used).36,37
An electrolyte redox couple is needed, usually I- /I3- or Co(II)/Co(III), to reduce the oxidised
dye remaining after electron injection into the semiconductor. Electron transfer from the
electrolyte to the dye prevents the recombination of the electron injected into the semiconductor
with the oxidised dye, and the redox couple receives an electron from the counter electrode,
which is supplied from an external load, to complete the circuit (Figure 8). In an optimised
DSSC electron injection and transport is faster than the recombination processes thus
minimising losses through recombination pathways (Figure 8).
Sensitisation of ZnO was initially carried out in the 1970s but the big breakthrough in the field
was made in 1991 when O’Regan and Grätzel published work on sensitised nanospheres of
TiO2 with a Ru dye to produce a DSSC with an efficiency of 7.1%.38 This was a significantly
higher efficiency than had previously been reported and is attributed to the increase in surface
area, and hence dye loading, from using TiO2 nanoparticles. Between 1991 and 2017 lab based
efficiencies increased steadily to just over 13%.39,40 Many Ru dyes with various ligands have
been investigated along with several other classes of dyes such as: triarylamines,41,42
squaraines,43 tiophenes,44 indolines,45 coumarins,46 porphyrins50 and pthalocyanines47 in an
10
effort to improve the light harvesting efficiency of devices. Generally, organic dyes have higher
extinction coefficients than Ru dyes, and can offer better response in the near-infrared region,
but have narrower absorption bands and can have problems with aggregation.44 The
development of co-sensitisation (the use of more than one dye) allows the absorption over
complementary regions of the visible spectrum to increase the overall light harvesting
efficiency of devices.39,40,48,49 Co-sensitisation, together with the use of cobalt electrolytes, has
led to some of the highest performing DSSC devices.39,40,50 For example the use a zinc
porphyrin (YD2-o-C8) co-sensitised with organic dye (Y123) yielded 12.3% efficiency under
standard conditions (AM1.5) and 13% at low light level conditions.50
Figure 8. Structure of a typical dye-sensitised solar cell (left) and electron transfer processes within a
device (right), desirable pathways are shown in green while recombination pathways are shown in
red.
The first publication on organohalide lead perovskite materials for solar cells, where the dye
component of a DSSC is replaced with a perovskite material, was in 2009.51 (The term
perovskite refers to particular crystal structure of general formula ABX3.) The perovskite
materials reported were methylammonium lead bromide and methylammonium lead iodide
producing efficiencies of 3.1 and 3.8%, respectively,51 however, these devices were fairly
unstable as the electrolyte dissolved the perovskite material. Two years later Park et al.
replaced the liquid electrolyte with a solid hole transport material producing a cell with 6.5%
efficiency and Snaith et al. achieved an efficiency of close to 11% using an alumina scaffold.
The device efficiencies have continued to rise, with the current certified record efficiency of
~22%.52,53
The work using alumina produced a step change in the field, alumina is an insulator and thus
charge injection is not energetically possible at this interface, showing that the perovskite
material itself has sufficient charge transport properties to function efficiently (more akin to a
thin-film solar cell). This widened the research field increasing the potential device
architectures that could successfully be employed.54–56 The current major challenge is the
stability of the perovskite material which is susceptible to degradation by moisture, and the
combination of light and oxygen.57 To overcome this, a range of mixed-cation and mixed-anion
perovskites have been developed that have improved stability, changing he halides can also
change the band gap (e.g. replacing some of the iodide anions with smaller bromide anions)
allowing the colour to be tuned.58–60 There are also environmental concerns over the use of lead
that will need to be addressed for successful commercialisation, possibly through the use of
encapsulation or lead replacement. The current situation is that, despite the concerns, the high
efficiencies, which are on par with market dominating silicon PVs, but at a potential fraction
of the monetary and energy cost, have drawn unparalleled research interest.
11
2.5.2. Organic Solar Cells
Organic solar cells (OSC) are devices, with a very thin film of a polymer or organic molecule
between two conductive electrodes. Light absorption initiates a series of very fast charge
transfer processes that lead to the formation of electrical charges (Figure 9).
The field of organic PVs started in 1986, when Tang et al. described the first operational device
consisting of a double-layered structure of copper phthalocyanine and a perylene derivative
(Figure 5C and D, respectively).61 This device demonstrated that solid organic electron donoracceptor interfaces can undergo charge transfer and then generate electrical current. Although
not efficient, this first device sparked widespread interest in organic PVs, which holds to this
day. OSC offer the numerous advantages over other PVs, including cheap production via
printing, the possibility to make the devices on flexible materials/substrates, and their efficient
operation at low light levels, making possible their use for indoor applications.
The most efficient devices are not simple bilayer structures, but are based on bulkheterojunction films with finely mixed donor and acceptor materials. The materials are usually
conjugated polymers, small molecules and fullerenes. Many new donors and acceptors have
been developed in the past ten years providing much better control over light harvesting and
charge transfer at the donor-acceptor interface. This has resulted in very reasonable device
efficiencies of ~12% keeping this technology attractive for researchers and
commercialisation.62–64 However, a further boost in efficiency and long terms stability are
required to make organic solar cells competitive with silicon or perovskite PVs.
Figure 9. Illustration of the structure of typical polymer-fullerene organic photovoltaic device. It
depicts the charge transfer process in the device and the movement of the electrical charges towards
the electrodes. Taken with permission from ref.65.
2.6.Probes and Chemosensors
The emission of light by molecules as either fluorescence or phosphorescence provides the
basis of highly sensitive analytical methods for studying and imaging chemical and biological
systems. The sensitivity is such that, under appropriate conditions, it is possible to detect close
to the level of single molecules. Stokes was probably the first person to demonstrate the
potential of this technique when he showed that light is emitted at longer wavelengths than
12
where it is absorbed.66 This difference, normally referred to as the Stokes shift, has many
practical applications, such as in fluorescence brightening agents, which are included in
detergents, paper and other commercial products, where they absorb light in the ultraviolet
region of the spectrum and emit in the blue to make things seem “whiter than white”. You have
probably seen the effect when white clothes appear blue upon irradiating with ultraviolet light.
It is also fundamental for providing the high sensitivity used in the analytical applications of
luminescence techniques.
Optical sensors and probes may involve either the absorption or emission of light. This has
been described in detail elsewhere.67 In this section we will concentrate on photoluminescence
(PL) sensors, in which changes in the intensity, lifetime or other properties of light emitted by
one molecular system are produced by the interaction with a second species, the analyte. One
typical case involves “quenching”, where the second molecule reduces the intensity of light
emitted by the first. The excited-state of the first molecule (M*) is produced by absorption of
light, and can either decay by light emission, heat loss, or “quenching” by the analyte (Q).
M* M + h
M* M +
M* + Q M + Q
Light emission (photoluminescence)
Heat loss (radiationless decay)
Quenching
The quenching competes with the other two processes in a way which depends on the
concentration of the analyte. This will lead to a decrease in both the intensity and lifetime of
the PL, and by measuring either of these we can have a direct measure of the concentration of
Q. This is frequently expressed mathematically as the Stern-Volmer equation, which expresses
the intensity (I, I0) or lifetime (, 0) of fluorescence as a linear function of the concentration
of concentration of Q. Ksv is the so-called Stern-Volmer constant, and the subscript 0 refers to
the property in the absence of quencher.
I0/I = 0/ = 1 + Ksv[Q]
Measurements of either fluorescence intensity or lifetime can, then, directly provide
information on quencher concentration. Such luminescence sensors have considerable
advantages over other techniques; they are cheap, respond rapidly to changes in concentration,
do not require any electrical contacts, and can easily be miniaturised. Since the field of PL
sensors is huge, we will limit ourselves to three important examples in which charge or electron
transfer are involved.
The applications of the metal complex Ru(bipy)32+ (Figure 5B) in solar energy conversion
through water splitting or DSSCs have been discussed earlier in this article. The same complex
has a very nice light emission in the visible region. This and closely related compound s are
widely used as PL sensors for oxygen.68,69 Emission is from the triplet metal to ligand charge
transfer (MLCT) state, which has a sufficiently long lifetime to readily be quenched by a variety
of species through energy or electron transfer. One of the most important examples is molecular
oxygen, and this is the basis for luminescent oxygen sensors, which are finding applications in
areas as diverse as blood gas analysis, oceanographic deep sea oxygen measurements,
combustion, food product quality control, and pressure sensitive paints for studying the
behaviour of cars and planes in wind tunnels. The metal complexes can be incorporated into a
wide variety of supports, including synthetic polymers, siloxanes and zeolites. These are
13
chosen to optimise efficiency under the working conditions, which may in some cases involve
extremes of temperature or pressure. These luminescence-based sensors, sometimes called
optrodes, are now one of the most convenient ways of measuring oxygen concentration in
solution.
Another area where PL quenching based on charge transfer is of great practical importance is
in explosives sensing. Current global political instabilities and the increasing awareness of
terrorist attacks are driving the development of more sensitive techniques for the detection of
explosive materials. Nitroaromatic compounds are a very important group of these, and are
used as both commercial and military explosives.70 One particularly well known examples is
2,4,6-trinitrotoluene (TNT, Figure 10A). Nitroaromatic compounds have a very high electron
affinity, i.e. a tendency to attract electrons, aromatic molecules in their excited-states are good
electron donors. If the aromatic molecule is also luminescent, interaction with the
nitroaromatics will lead to quenching, providing a good way of detecting and quantifying TNT
or other explosives. The sensitivity can be enhanced still further through fluorescence
quenching using conjugated polymer-based PL sensors71,72 Conjugated polymers are formed
by joining together a series of -conjugated systems, such as benzene, acetylene, aniline and
thiophene to form highly conjugated systems in which the electrons are delocalised over the
whole molecule. They behave as organic semiconductors, and examples are given in Figure
10C-F. In some cases, these are luminescent. Such systems are now widely used in large area
flat screen light emitting diode televisions, computer monitors and other displays. However,
they are also becoming increasingly important as light emitters in various optical sensors and
probes. When the polymer is excited by absorption of light, the excitation is delocalised along
the polymer chain as what is termed an exciton. Emission can be from anywhere along the
chain, and these are termed amplifying fluorescent polymers. This leads to an enhancement of
the fluorescence quenching by analytes, such as TNT, which makes these excellent candidates
for highly sensitive explosive sensors. An example of the principle of superquenching is shown
in Figure 10C where the fluorescence of the conjugated polymer poly(9,9-dioctylfluorene-2,7diyl]-co-bisthiophene] (PF2T) is quenched by dinitrobenzene. This forms the basis of a
handheld device for the detection of explosives.73
14
A
B
CH3
O2N
NO2
S
H3C(H2C)6H2C
S
CH2(CH2)6CH3
n
NO2
D
C
n
n
F
E
X
H
N
X
H
N
n
N
H
(X= NH, S)
G
n
H
N
N
N
N
Zn
Cl
Cl
N
Cl
Zn
Cl
N
N
N
Figure 10. Structures of (A) 2,4,6-trinitrotoluene (TNT) and examples of conjugated polymers: (B)
poly(9,9-dioctylfluorene-2,7-diyl]-co-bisthiophene (PF2T), (C) poly(acetylene); (D) poly(p-phenylene
vinylene); (E) poly(pyrrole) (X= NH) and poly(thiophene) (X=S), (F) poly(aniline), (G) anthracene
based chemosensor and (H) its zinc complex.
Intensity (a.u)
1.0
0.8
0.6
0.4
0.2
0.0
500
600
700
800
Wavelength (nm)
Figure 11. Fluorescence spectrum of PF2T in ethylcellulose films alone (solid line) and exposed to
dinitrobenzene (dashed line).73
As a final example, we choose a system where interaction of the analyte with the sensor leads
to an increase in fluorescence intensity. The anthracene derivative in Figure 10G has two
substituted ethylenediamine groups chemically bonded to the aromatic moiety, and, unlike
many anthracene derivatives, is not fluorescent. Amine groups are known to be capable of
15
transferring electrons to excited aromatic molecules, and electron or charge transfer quenches
the normal fluorescence of the anthracene. However, ethylenediamine groups also bind to
metal ions, such as zinc(II), very strongly to form chelate complexes. In contrast to compound
A, the zinc complex in Figure 10H is strongly fluorescent (see emission spectra in Figure 11).74
This complexation “switches on” the fluorescence, and this provides the possibility of probing
metal ions such as zinc and calcium, which is particularly valuable for studying them in
biological systems. Compounds such as this, containing a fluorescent unit (the fluorophore), a
binding site, and a mechanism by which one affects the other are termed “fluorescent”
chemosensors. A large number of this have been prepared, and are valuable for sensing metal
ions, protons (H+), anions, and various organic and biological targets.75
2.7.Optimisation of device efficiencies; triplet or singlet state photochemistry?
In this article we have concentrated on the general charge transfer features of the technologies
described, rather than the detailed photochemistry, but one important aspect of photochemical
behaviour is worth discussing because it is a very active area of current research in device
design; that is the difference in the properties of triplet and singlet excited states.
Figure 12. The Jablonski Diagram. Radiative transitions are indicated by straight line arrows
whereas non-radiative transitions are represented by wavy line arrows. The range of rate constants
generally observed is given in parentheses
Triplet excited-states in molecules and polymers are formed via intersystem crossing (Figure
12), which is a radiationless process from a singlet to a triplet state involving a spin flip of one
of the electrons.76 Triplets are different from singlets in their spin multiplicity, which makes
the transition between them very slow, sometimes taking seconds; hence, most triplet state
lifetimes are milliseconds (ms) and even seconds. The formation of triplets opens a
photochemical degradation channel in conjugated polymers and molecules, used for electronic
16
applications, because they make possible energy transfer to molecular oxygen (which has a
triplet ground-state), thus promoting oxygen to its highly reactive singlet state.77 There are
other mechanisms of triplet state formation including electron recombination at donor-acceptor
interfaces and singlet fission.78,79 The former is a common process in organic-light emitting
diodes, used nowadays in displays. In such devices, electrons are injected into an organic film,
where they encounter each other and form emissive excited-states via back electron transfer.
Due to the lack of spin correlation between electrons, three in every four of these encounters
generate non-emissive triplet excitons, which has implications for design and device stability.78
Singlet fission requires close interaction between two chromophores with a triplet state energy
half of the singlet.79 It leads to the formation of two triplet states from one photoexcited singlet.
This process has a high potential for application in PVs, where two electrons can be generated
from a single absorbed photon, thus doubling electrical current, although reducing
photovoltage.80 Optimisation of charge transfer from these triplets is part of the implementation
of this process for actual devices; it has recently been demonstrated to be highly efficient for
pentacene-CdSe interfaces, but further work is required for its utilisation in PVs.81
3. Summary and Outlook
Photoinduced charge transfer is the key chemical reaction in many technologies. In 1917
science lagged behind application, imaging technologies using these reactions were already
well established but the fundamental science was little understood. Over the past one hundred
years, developments in chemical applications of quantum mechanics have provided a deep
understanding of molecular and semiconductor structure and the interaction of light and matter.
Developments in the application of photoinduced charge transfer have led to new technologies
e.g. the various types of solar cells, photocatalysts for water purification, and devices for ultrasensitive chemical sensing discussed here. We do not know what Science Progress will give
us over the next one hundred years, but there can be little doubt that photoinduced charge
transfer will continue to play an important role in the technologies of the future.
4. Acknowledgments
HDB is grateful for funding from “The Coimbra Chemistry Centre” which is supported by the
Fundação para a Ciência e a Tecnologia (FCT), Portuguese Agency for Scientific Research,
through the programmes UID/QUI/UI0313/2013 and COMPETE. MLD is grateful for the
financial support of the EPSRC and Innovate UK for the SPECIFIC Innovation and Knowledge
Centre (grant numbers EP/I019278/1, EP/K000292/1, EP/L010372/1) and the European
Regional Development Fund through the Welsh Government for support to the Sêr Solar
program (MLD and CDC). This work is part-funded by the European Regional Development
Fund through the Welsh Government (SD).
5. References
1
2
3
4
5
N. Bohr, Philos. Mag. Ser. 6, 1913, 26, 857–875.
W. Heisenberg, Zeitschrift für Phys., 1925, 33, 879–893.
E. Schrödinger, Ann. Phys., 1926, 384, 361–376.
H. Gernscheim, The Origins of Photography, Thames & Hudson, London, 1982.
E. Becquerel, Comptes Rendus, 1839, 9, 561–567.
17
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
S.-H. Bae, H. Zhao, Y.-T. Hsieh, L. Zuo, N. De Marco, Y. S. Rim, G. Li and Y. Yang,
Chem, 2016, 1, 197–219.
A. M. Kuznet︠s︡ov, Charge transfer in chemical reaction kinetics, Presses polytechniques
et universitaires romandes, Lausanne, 1997.
P. Ceroni and V. Balzani, in The Exploration of Supramolecular Systems and
Nanostructures by Photochemical Techniques, ed. P. Ceroni, Springer, 2012.
P. Suppan, Chemistry and Light, The Royal Society of Chemistry, 1994.
L. J. Schaaf, Out of the Shadows, Herschel, Talbot, and the Invention of Photography,
New Haven and London, Yale, 1992.
H. Gernsheim and A. Gernsheim, Concise History of Photography, Dover Publications,
New York, 3rd Ed., 1986.
W. Crawford, The Keepers of Light: A History & Working Guide to Early Photographic
Processes, Morgan and Morgan, New York, 1979.
J. Cat, Maxwell, Sutton, and the Birth of Colour Photography, a Binocular Study,
Palgrave Macmillan, New York, 2013.
R. M. Evans, Sci. Am., 1961, 205, 118–128.
W. West and P. B. Gilman, in The Theory of the Photographic Process, ed. T. H. James,
MacMillan Pub Co, New York, 4th edn., 1977.
H. Vogel, Berichte der Dtsch. Chem. Gesellschaft, 1873, 6, 1302–1306.
A. Horder, The Ilford Manual of Photography, Ilford Ltd, Essex, 1958.
F. J. Hurley, J. Wright and F. Hurley, South with ‘Endurance’: Antarctic Photographs,
Bloomsbury Publishing, London, 2001.
D. J. Mitchell, Notes Rec. R. Soc. Lond., 2010, 64, 319–337.
R. W. Gurney and N. F. Mott, Proc. R. Soc. London Ser. A, 1938, 146, 151.
R. H. Bube, Photoelectronic Properties of Semiconductors, Cambridge University
Press, Cambridge, 1992.
G. B. Evans, M. B. Ledger and H. H. Adam, in Applied Photochemistry, eds. R. C.
Evans, P. Douglas and H. D. Burrows, Springer, 2013, pp. 363–402.
T. Tani, Ed., Photographic Science, Oxford University Press, Oxford, 2011.
T. H. James, Ed., The Theory of the Photographic Process, MacMillan Pub Co, New
York, 4th edn.
C. G. Hatchard and C. A. Parker, Proc. R. Soc. London. Ser. A. Math. Phys. Sci., 1956,
235, 518 LP-536.
D. F. Zigler, E. C. Ding, L. E. Jarocha, R. R. Khatmullin, V. M. DiPasquale, R. B. Sykes,
V. F. Tarasov and M. D. E. Forbes, Photochem. Photobiol. Sci., 2014, 13, 1804–1811.
M. Montalti, A. Credi, L. Prodi and M. T. Gandolfi, Handbook of Photochemistry,
Taylor and Francis, 2006.
J. F. W. Herschel, Phil. Trans. R. Soc., 1842, 132, 181–214.
J. O. Bockris, in Comprehensive Treatise of Electrochemistry: Volume 3:
Electrochemical Energy Conversion and Storage, eds. J. O. Bockris, B. E. Conway, E.
Yeager and R. E. White, Springer US, Boston, MA, 1981, pp. 505–526.
P. Tseng, J. Lee and P. Friley, Energy, 2005, 30, 2703–2720.
J. R. Darwent, P. Douglas, A. Harriman, G. Porter and M.-C. Richoux, Coord. Chem.
Rev., 1982, 44, 83–126.
N. Sutin and C. Creutz, Pure Appl. Chem., 1980, 52, 2717—2738.
A. Fujishima and K. Honda, Nature, 1972, 238, 37–38.
Y. Nosaka, A. Y. Nosaka, W. S. Jenks, C. Minero, V. Maurino, D. Vione, B. Ohtani, N.
Shaham-Waldmann, Y. Paz, D. Robert, V. Keller, N. Keller, M. Matsuoka, T. Toyao,
Y. Horiuchi, M. Takeuchi, M. Anpo, Z. Wang, C. Chen, W. Ma, J. Zhao, R. Amadelli,
L. Samiolo, S. Liu, M. Lim, R. Amal, J. A. Rengifo-Herrera, A. G. Rincón, C. Pulgarin,
18
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
D. Ollis, O. M. Alfano, A. E. Cassano, R. J. Brandi, M. L. Satuf, T. Ochiai, A. Fujishima,
S. Malato, P. Fernández-Ibáñez, M. I. Maldonado, I. Oller and M. I. Polo-López,
Photocatalysis and Water Purification: From Fundamentals to Recent Applications,
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, Wiley-VCH., 2013.
A. Fujishima, X. Zhang and D. A. Tryk, Surf. Sci. Rep., 2008, 63, 515–582.
A. Hagfeldt, G. Boschloo, L. Sun, L. Kloo and H. Pettersson, Chem. Rev., 2010, 110,
6595–6663.
M. Grätzel, J. Photochem. Photobiol. C Photochem. Rev., 2003, 4, 145–153.
B. O’Regan and M. Gratzel, Nature, 1991, 353, 737–740.
S. Mathew, A. Yella, P. Gao, R. Humphry-Baker, C. F. E., N. Ashari-Astani, I.
Tavernelli, U. Rothlisberger, N. Khaja and M. Grätzel, Nat Chem, 2014, 6, 242–247.
K. Kakiage, Y. Aoyama, T. Yano, K. Oya, J. Fujisawa and M. Hanaya, Chem. Commun.,
2015, 51, 15894–15897.
P. J. Holliman, M. Mohsen, A. Connell, M. L. Davies, K. Al-Salihi, M. B. Pitak, G. J.
Tizzard, S. J. Coles, R. W. Harrington, W. Clegg, C. Serpa, O. H. Fontes, C.
Charbonneau and M. J. Carnie, J. Mater. Chem., 2012, 22, 13318–13327.
A. Connell, P. J. Holliman, E. W. Jones, L. Furnell, C. Kershaw, M. L. Davies, C. D.
Gwenin, M. B. Pitak, S. J. Coles and G. Cooke, J. Mater. Chem. A, 2015, 3, 2883–2894.
A. Connell, P. J. Holliman, M. L. Davies, C. D. Gwenin, S. Weiss, M. B. Pitak, P. N.
Horton, S. J. Coles and G. Cooke, J. Mater. Chem. A, 2014, 2, 4055.
S. M. Abdalhadi, A. Connell, X. Zhang, A. A. Wiles, M. L. Davies, P. J. Holliman and
G. Cooke, J. Mater. Chem. A, 2016, 4, 15655–15661.
S. Ito, S. M. Zakeeruddin, R. Humphry-Baker, P. Liska, R. Charvet, P. Comte, M. K.
Nazeeruddin, P. Péchy, M. Takata, H. Miura, S. Uchida and M. Grätzel, Adv. Mater.,
2006, 18, 1202–1205.
K. Hara, Z.-S. Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y. Dan-oh, C. Kasada,
A. Shinpo and S. Suga, J. Phys. Chem. B, 2005, 109, 15476–15482.
J.-H. Yum, S. Jang, R. Humphry-Baker, M. Grätzel, J.-J. Cid, T. Torres and M. K.
Nazeeruddin, Langmuir, 2008, 24, 5636–5640.
P. J. Holliman, K. J. Al-Salihi, A. Connell, M. L. Davies, E. W. Jones and D. A. Worsley,
RSC Adv., 2014, 4, 2515–2522.
M. L. Davies, T. M. Watson, P. J. Holliman, A. Connell and D. A. Worsley, Chem.
Commun. (Camb)., 2014, 50, 12512–4.
A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W.-G.
Diau, C.-Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science (80-. )., 2011, 334, 629–
634.
A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Soc., 2009, 131, 6050–
6051.
M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami and H. J. Snaith, Science (80-. ).,
2012, 338, 643–647.
National Renewable Energy Laboratory (NREL) Home Page. Retrieved from
http://www.nrel.gov/
M. J. Carnie, C. Charbonneau, M. L. Davies, J. Troughton, T. M. Watson, K.
Wojciechowski, H. Snaith and D. A. Worsley, Chem. Commun., 2013, 49, 7893–7895.
H.-S. Kim, S. H. Im and N.-G. Park, J. Phys. Chem. C, 2014, 118, 5615–5625.
M. J. Carnie, C. Charbonneau, M. L. Davies, B. O. Regan, D. A. Worsley and T. M.
Watson, J. Mater. Chem. A, 2014, 2, 17077–17084.
N. Aristidou, I. Sanchez-Molina, T. Chotchuangchutchaval, M. Brown, L. Martinez, T.
Rath and S. a. Haque, Angew. Chemie - Int. Ed., 2015, 54, 8208–8212.
M. Saliba, T. Matsui, J.-Y. Seo, K. Domanski, J.-P. Correa-Baena, N. Mohammad K.,
19
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
S. M. Zakeeruddin, W. Tress, A. Abate, A. Hagfeldt and M. Gratzel, Energy Environ.
Sci., 2016, 9, 1853–2160.
J. H. Noh, S. H. Im, J. H. Heo, T. N. Mandal and S. Il Seok, Nano Lett., 2013, 13, 1764–
1769.
M. L. Davies, M. Carnie, P. J. Holliman, A. Connell, P. Douglas, T. Watson, C.
Charbonneau, J. Troughton and D. Worsley, Mater. Res. Innov., 2014, 18, 482–485.
C. W. Tang, Appl. Phys. Lett., 1986, 48, 183–185.
S. Li, L. Ye, W. Zhao, S. Zhang, S. Mukherjee, H. Ade and J. Hou, Adv. Mater., 2016,
28, 9423–9429.
D. Baran, R. S. Ashraf, D. A. Hanifi, M. Abdelsamie, N. Gasparini, J. A. Rohr, S.
Holliday, A. Wadsworth, S. Lockett, M. Neophytou, C. J. M. Emmott, J. Nelson, C. J.
Brabec, A. Amassian, A. Salleo, T. Kirchartz, J. R. Durrant and I. McCulloch, Nat
Mater, 2016, advance on.
K. Ardani and R. Margolis, Nrel, 2011, 1–136.
S. D. Dimitrov and J. R. Durrant, Chem. Mater., 2014, 26, 616–630.
G. G. Stokes, Philos. Trans. R. Soc. London , 1852, 142, 463–562.
R. C. Evans and P. Douglas, in Applied Photochemistry, eds. R. C. Evans, P. Douglas
and H. D. Burrow, Springer Netherlands, Dordrecht, 2013, pp. 403–434.
V. M. Chauhan, F. Giuntini and J. W. Aylott, Sens. Bio-Sensing Res., 2016, 8, 36–42.
J. N. Demas and B. A. DeGraff, J. Chem. Educ., 1997, 74, 690.
J. Akhavan, The Chemistry of Explosives, The Royal Society of Chemistry, 2011.
D. T. McQuade, A. E. Pullen and T. M. Swager, Chem. Rev., 2000, 100, 2537–2574.
S. W. Thomas, G. D. Joly and T. M. Swager, Chem. Rev., 2007, 107, 1339–1386.
T. Neves, L. Marques, L. Martelo and H. D. Burrows, Proc. IEEE Sensors, 2014, 1866.
A. W. Czarnik, Acc. Chem. Res., 1994, 27, 302–308.
J. F. Callan, A. P. de Silva and D. C. Magri, Tetrahedron, 2005, 61, 8551–8588.
J. D. Kovac, J. Chem. Educ., 1998, 75, 545.
S. Chambon, A. Rivaton, J.-L. Gardette, M. Firon and L. Lutsen, J. Polym. Sci. Part A
Polym. Chem., 2007, 45, 317–331.
R. H. Friend, R. W. Gymer, A. B. Holmes, J. H. Burroughes, R. N. Marks, C. Taliani,
D. D. C. Bradley, D. A. Dos Santos, J. L. Bredas, M. Logdlund and W. R. Salaneck,
Nature, 1999, 397, 121–128.
M. B. Smith and J. Michl, Chem. Rev., 2010, 110, 6891–6936.
D. N. Congreve, J. Lee, N. J. Thompson, E. Hontz and S. R. Yost, 2007.
B. Ehrler, B. J. Walker, M. L. Böhm, M. W. B. Wilson, Y. Vaynzof, R. H. Friend and
N. C. Greenham, Nat. Commun., 2012, 3, 1019.
20