You are on page 1of 19

International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

www.elsevier.com/locate/ijhydene

Efficiency of solar water splitting using semiconductor electrodes


A.B. Murphya, c , P.R.F. Barnesa, c , L.K. Randeniyaa, c , I.C. Plumba, c,∗ , I.E. Greyb, c ,
M.D. Horneb, c , J.A. Glasscocka, c
a CSIRO Industrial Physics, P.O. Box 218, Lindfield NSW 2070, Australia
b CSIRO Minerals, Box 312, Clayton South VIC 3169, Australia
c CSIRO Energy Transformed Flagship, Australia

Available online 13 March 2006

Abstract
Reliable measurement of the photoconversion efficiency for semiconductor electrodes is essential to the assessment of electrode
performance. In this paper, the influence of the choice of light source on measured photoconversion efficiencies for semiconductor
photoelectrodes is examined. Measurements of efficiency performed under xenon lamp and solar illumination are compared with
efficiencies calculated by integrating the incident photon conversion efficiency (IPCE) over the lamp and solar spectra. It is shown
that use of a xenon lamp as the light source can lead to a large overestimate of the photoconversion efficiency, relative to that
obtained under standard AM1.5 solar illumination. The overestimate is greater when a water filter is fitted to the xenon lamp, and
when a wide-band gap semiconductor such as TiO2 is used as the photoelectrode. Achievable photoconversion efficiencies using
rutile TiO2 are calculated taking into account the losses due to imperfefct absorption, reflection and charge-carrier recombination;
these calculated efficiencies agree with the measurements to within experimental uncertainties. It is demonstrated that many
photoconversion efficiencies presented in the literature are overestimated. It is concluded that reliable estimation of efficiency
under standard conditions is best obtained by measuring the IPCE as a function of wavelength, and integrating over the AM1.5
solar spectrum, or by measuring under sunlight with a similar zenith angle to that of the AM1.5 spectrum.
䉷 2006 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.

Keywords: Hydrogen generation; Solar energy; Photoelectrochemistry; Photoelectrolysis; Photocatalysis; Spectral irradiance; Xenon lamp;
Solar spectrum; Semiconducting materials; Titanium dioxide

1. Introduction ratio of the chemical potential energy stored in the


form of hydrogen molecules to the incident radiative
Photoelectrochemical splitting of water to produce energy. The benchmark efficiency is 10%, which is
gaseous hydrogen using solar energy has been re- generally considered to be required for commercial
searched intensively since its initial demonstration in implementation [2].
1972 [1]. The most important figure of merit for a Most measurements of the photoconversion effi-
semiconductor photoelectrode is the photoconversion ciency are performed under illumination by artificial
efficiency for water splitting, which is defined as the light sources. This is convenient for many reasons—the
artificial sources are stationary, and their intensity is
∗ Corresponding author. CSIRO Industrial Physics, P.O. Box 218, essentially constant with time, while the spectrum and
Lindfield NSW 2070, Australia. Tel.: +61 2 9413 7351; intensity of solar radiation reaching the ground depends
fax: +61 2 9413 7631. on the time of day, atmospheric conditions such as
E-mail address: ian.plumb@csiro.au (I.C. Plumb). cloud cover, water vapour content and ozone column,
0360-3199/$30.00 䉷 2006 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2006.01.014
2000 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

albedo of the surrounding ground, etc. However, the radiation. These include the band gap of the semicon-
comparison of photoconversion efficiencies obtained for ductor, reflection of radiation before reaching the semi-
different semiconductor materials requires that these ef- conductor, the level of absorption of the radiation in
ficiencies are presented for a standard solar spectrum, the semiconductor, and the transport of charge carriers
usually the AM1.5 global solar spectrum, and artificial through the semiconductor. We examine the influence of
light sources do not accurately replicate such spectra. each of these factors on the achievable efficiency, both
An examination of the literature reveals that in most for solar illumination and xenon lamp illumination.
cases the photoconversion efficiency values presented We have measured the photoconversion efficiency
are obtained using illumination by a particular artificial for a cell with a rutile TiO2 photoelectrode under both
light source. For example, Nozik [3] used a UV source xenon lamp and solar illumination. We have also mea-
with radiation between 300 and 400 nm to measure pho- sured the incident photon conversion efficiency (IPCE),
toconversion efficiencies of up to 4% for TiO2 photo- which is the efficiency of conversion of photons inci-
electrodes. Giordano et al. [4] obtained an efficiency of dent on the photoelectrochemical cell to photocurrent
2.7% for a TiO2 electrode doped with platinum using flowing between the working and counter electrodes, as
a mercury lamp. Khan and Akikusa [5] quoted an effi- a function of wavelength of the incident radiation. This
ciency for flame oxidised titanium of 2.0%; the publi- allows the photoconversion efficiency to be calculated
cation indicates a xenon lamp was used, but Akikusa’s for any radiation spectrum. The efficiencies obtained
PhD thesis [6] reveals that a mercury–xenon lamp was from the measurements are compared with those calcu-
used. Akikusa and Khan [7] reported an efficiency of lated taking into account the spectrum of the radiation,
1.6% for a similar photoelectrode under xenon lamp the band-gap, the absorption and reflection of the radi-
illumination. Khan et al. [8] claimed an efficiency of ation, and transport of the charge carriers.
8.35% for a flame-pyrolysed titanium photoelectrode The measurements reported here were obtained at
with a band-gap extended to 535 nm by carbon doping; room temperature (T = 295 K). The effect of increased
again xenon lamp illumination was used. Mishra et al. temperatures on water splitting efficiencies has been
[9] obtained photoconversion efficiencies as high as 3% investigated by Licht [12,13].
for TiO2 photolectrodes using a mercury–xenon lamp In Section 2, we present the details of our mea-
source. Tang et al. [10] reported an efficiency of 2.5% surements, including the characterisation of the light
for a mesoporous anatase photoelectrode using a xenon sources, and the results of these measurements. Our cal-
lamp. Radecka et al. [11] obtained an efficiency of 1.9% culations of the efficiency limits for the different light
using a xenon lamp for a mixed anatase and rutile thin sources are given in Section 3. The measurements are
film deposited by radio-frequency sputtering. compared with the calculations in Section 4; in addition,
As will be shown later, all of these efficiency figures the results of other workers are discussed in the light of
are close to or higher than those that are thermody- our results, and possible means of increasing the pho-
namically possible for the standard AM1.5 global so- toconversion efficiency of semiconductor electrodes are
lar spectrum, given the band gap of the semiconductor considered. Conclusions are presented in Section 5.
material used and the bias voltage applied. It therefore
seems likely that the efficiencies are larger than those
that would be obtained for sunlight. 2. Experimental details and results
In this study, we investigate the influence of the light
spectrum on the photoconversion efficiency for cells 2.1. Sample preparation
with semiconductor photoelectrodes. We consider in
particular the most widely used source, the xenon arc The specimen used in this study was prepared by
lamp. We have had the spectrum of our lamp charac- oxidising a titanium substrate in a methane and oxy-
terised by an accredited standards laboratory. We exam- gen flame. Before oxidation, the titanium sheet (Sigma
ine the influence on efficiency of the spectrum of the Aldrich, 99.7%, 0.25 mm thick) was polished and
lamp, both with and without a water filter fitted. Water then etched in Kroll’s solution (one part 40% HF,
filters are widely used to absorb the infrared output of one part 70% HNO3 and three parts water) for 5 s. A
the lamp, thereby decreasing the heating of the irradi- small (6 mm × 6 mm) piece of the sheet was held in a
ated surfaces. We also investigate the influence of the methane–oxygen flame for five minutes, at a position
age of the lamp tube on the spectrum. at which the flame temperature was measured to be
The photoconversion efficiency depends on many 850 ◦ C using a K-type thermocouple. The oxygen to
factors in addition to the spectrum of the incident methane volume ratio of the gas mixture was 1.15, so
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2001

0.35

Diffuse reflectance
0.30

0.25

0.20

0.15

300 400 500 600 700 800


Specular reflectance

0.05
Measurement
0.04 Best fit, h = 280 nm
0.03

0.02

0.01

0.00
300 400 500 600 700 800
Wavelength (nm)

Fig. 1. Diffuse and specular reflectance of the rutile TiO2 film formed by flame-oxidising a titanium substrate. The reflectance is measured
relative to a matt Teflon standard. The best fit to the specular reflectance (assuming 15% by volume of the TiO2 film is air) is also shown. The
specular reflectance fit is performed only for wavelengths greater than 400 nm, since the strong absorption at shorter wavelengths precludes
the formation of an interference pattern.

the flame was lean with respect to stoichiometry. An The thickness of the oxide layer was derived from the
oxide layer was formed on both sides of the titanium interference pattern evident in the specular reflectance
substrate. The side of the substrate exposed directly spectrum. This was done using a least squares fit of the
to the flame is analysed here. The colour of the oxide specular reflectance from 400 to 800 nm to the expres-
layer was mid-grey, with some red and green colours sion [15] for the reflection coefficient of an absorbing
indicating the formation of an interference pattern. The film of a given thickness on an absorbing substrate. The
preparation and characterisation of a range of similar real part of the refractive index of rutile was taken from
samples have been described in more detail elsewhere Cardona and Harbeke [16] and DeVore [17], as reported
[14]. by Ribarsky [18] and the imaginary part from the work
of Eagles [19]. Scanning electron microscope images
2.2. Sample characterisation indicated that the sample had a porosity of 15 ± 10%.
The refractive index was accordingly reduced to allow
X-ray diffraction analysis of the sample was per- for 15 ± 10% by volume of air, using the formula given
formed using a Philips diffractometer with a Cu K by Mergel [20], based on the Bruggeman approxima-
source. The measurements indicated that the oxide layer tion. The best fit was obtained for TiO2 film thickness
had a rutile structure. h = 280 ± 40 nm, with the uncertainty allowing for the
The diffuse and total reflectance of the sample estimated range of air volumes. The best fit curve for
were measured as functions of wavelength from 250 15% air by volume is shown in Fig. 1. The feature of
to 800 nm using the diffuse reflectance attachment of the fitted curve that determines h is the separation of
a Cary 5 spectrophotometer. The reflectances were the peaks of the interference pattern. The amplitude
measured relative to a Teflon standard. The specular of the fitted curve is normalised to have the same stan-
reflectance was obtained by subtracting the diffuse dard deviation from the mean specular reflectance as
reflectance from the total reflectance. The diffuse and the measured curve.
specular reflectances are shown in Fig. 1. The diffuse
reflectance curve shows a strong decrease in the re- 2.3. Light sources
flectance at wavelengths below 400 nm, indicating an
absorption edge at around 3.1 eV, close to the band gap The spectrum of a light source is usually described
of rutile. by the spectral irradiance E () (units W m−2 nm−1 ),
2002 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

defined by E = dES /d, where The zenith and azimuthal angles of the sun were, re-
 ∞ spectively, 55.4◦ and 357.8◦ . The surface of the detec-
ES = E () d (1) tor was tilted to face the sun. The calculation uses these
0 parameters. Values corresponding to the AM1.5 stan-
is the incident irradiance. The incident spectral photon dard spectrum were used for all other parameters (e.g.,
flux I (units m−2 nm−1 s−1 ) is related to the spectral atmospheric composition and surface albedo).
irradiance by The calculated solar spectrum is similar to the AM1.5
global solar spectrum. The spectral irradiance of the
E ()
I () = , (2) calculated spectrum is slightly lower at ultraviolet
hc and visible wavelengths, and similar at infrared wave-
where h is Planck’s constant and c is the speed of light in lengths. The total irradiance of the calculated spectrum
a vacuum. The spectral photon flux is the more relevant was 962 W m−2 , consisting of 835 W m−2 direct and
quantity in water splitting applications, since at most 127 W m−2 diffuse irradiance. The respective figures
one electron–hole pair is produced per absorbed photon, for the AM1.5 global spectrum are 1000 W m−2 , con-
and photon energy above the required energy to split a sisting of 900 W m−2 direct and 100 W m−2 diffuse
water molecule is lost. irradiance.

2.3.1. AM1.5 global solar spectrum 2.3.3. Xenon arc lamp


The standard reference solar spectra for calculating We used an Oriel 6271 ozone-free xenon lamp, with
the conversion efficiency of solar energy to electrical an Oriel 61945 water filter fitted. The water filter was
or chemical energy are the AM1.5 spectra, defined by filled with deionised water, and removed infrared ra-
the American Society for Testing and Materials [21]. diation above about 1000 nm and ultraviolet radiation
The AM, or air mass, factor characterises the effect of below about 250 nm. An Oriel digital exposure control
the Earth’s atmosphere on the solar radiation, and is was used to maintain the total radiant power of the lamp
given by AM=1/ cos , where  is the solar zenith angle constant.
(the angle between the overhead and actual position of The spectral irradiance of the xenon lamp was mea-
the sun). sured by the Australian National Measurement Institute
Two AM1.5 spectra are defined. They describe re- [24]. Measurements were performed using a photomul-
spectively the direct, and the total, or global, spectral ir- tiplier detector for wavelengths from 240 to 400 nm, a
radiance at air mass 1.5 under standard atmospheric and silicon detector from 300 to 1010 nm, an indium gallium
surface conditions. AM1.5 corresponds to a solar zenith arsenide detector from 900 to 1680 nm, and a lead sul-
angle of 48.2◦ , at which angle the sun’s radiation has phide detector from 1300 to 2500 nm. The use of over-
to pass through 1.5 times the thickness of atmosphere lapping regions ensured that the measurements made
relative to a solar zenith angle of 0◦ . The receiving sur- using the different detectors were comparable. The mea-
face is inclined at 37◦ , so the incidence angle is 11.2◦ . surement intervals used were 2 nm for wavelengths be-
The global AM1.5 spectrum includes radiation incident low 400 nm, 5 nm for wavelengths between 400 and
from all angles, and therefore includes scattered light. 1200 and 10 nm for longer wavelengths.
The direct AM1.5 spectrum includes only the compo- Spectra were measured both with and without the wa-
nent of the global spectrum that is directly incident from ter filter fitted to the xenon lamp. While the water filter
the sun. The global spectrum is the relevant quantity was used in our efficiency measurements, calculations
for solar water splitting experiments, since both direct for both spectra are presented in order to broaden the
and scattered light will be incident on a water-splitting relevance of our results.
reactor. The spectrum of the lamp without the water filter ex-
tends to wavelengths longer than 2.5 m, which is the
2.3.2. Solar spectrum for efficiency measurement limit for which the measurements could be performed.
The spectrum of sunlight at the time and location of However, by measuring the total radiant power flux with
photoconversion efficiency measurement was calculated a thermopile, and comparing this to the power obtained
using the SMARTS code, version 2.9.2 [22,23]. (This is by integrating the lamp spectrum to 2.5 m, it was de-
the same code that has been used to derive the standard termined that less than 5% of the power was at wave-
AM1.5 spectra.) The measurement was made in Syd- lengths longer than 2.5 m.
ney, Australia (latitude 33.87◦ S, longitude 151.22◦ E) at Fig. 2 shows the spectral photon flux of the xenon
midday on 28 May 2004. There was no cloud cover. lamp with and without the water filter. Both spectra have
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2003

35 Xenon lamp, without water filter


Xenon lamp, with water filter

Spectral photon flux (1018 m-2 s-1 nm-1)


AM1.5 global sunlight
30
6

25 5

4
20
3

15 2

1
10
0
200 300 400 500 600 700
5

0
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400
Wavelength (nm)

Fig. 2. Spectral photon flux for the xenon lamp, with and without the water filter, compared to the AM1.5 global solar spectrum. The lamp
had logged 1250 h of operation when the spectrum was recorded. The inset shows detail of short wavelengths. In all cases, the total irradiance
is normalised to 1000 W m−2 . Measurements were performed by the National Measurement Institute.

been normalised to a total irradiance of 1000 W m−2 , of a replacement tube. These measurements were
to facilitate comparison with the AM1.5 global solar performed by measuring the lamp irradiance with a
spectrum. thermopile, after passing the light through a monochro-
The lamp spectra are almost identical for wavelengths mator with 12 nm pass band. The thermopile is
up to around 900 nm, apart from the influence of the described in Section 2.4. The measurements were per-
normalisation factor. The water filter absorbs a signif- formed with the water filter fitted to the lamp, and the
icant fraction of the light at wavelengths between 900 total irradiance was normalised to 1000 W m−2 in each
and 1100 nm, and all of the radiation at longer wave- case. To check the method of measurement, the spec-
lengths. In total, the filter absorbs about 50% of the lamp trum was also measured for the tube with 1250 h of op-
irradiance; hence when the spectra were normalised to eration; the results are almost identical to those shown
the same total irradiance, the spectral irradiance of the in Fig. 2 apart from the poorer wavelength resolution.
lamp without the filter is approximately 50% lower for The spectral photon flux for the three cases is shown
wavelengths less than 900 nm. in Fig. 3. The new tube has a greater proportion of the
Fig. 2 indicates that at short wavelengths, the spec- photon flux in the ultraviolet. Towards the end of tube
tral photon flux of the xenon arc lamp is greater than life, the proportion of the photon flux in the ultraviolet
that of AM1.5 global sunlight. The lamp emits a signifi- decreases rapidly. Unless otherwise noted, the results
cant photon flux for wavelengths from 250 nm, whereas presented in the remainder of the paper pertain to the
the solar photon flux is minimal at wavelengths below spectra measured for the tube with 1250 h of operation.
300 nm. The enhanced photon flux at short wavelengths Other lamps, such as non-ozone-free xenon arc
is greater for the lamp with the water filter, as a conse- lamps, mercury–xenon arc lamps, and mercury lamps,
quence of the normalisation. are also used for measurement of photoconversion ef-
The xenon lamp spectrum is similar to that presented ficiencies. Such lamps emit more strongly in the UV
by the manufacturer [25]. However, arc lamp spectra range than ozone-free xenon lamps [25].
vary from lamp to lamp, depending on the age of the
tube and other variables. The tube used here had logged 2.4. Measurement of efficiency
1250 h of operation before calibration. To investigate
the extent of the variability of the lamp spectrum, we Two light sources were used for the water-splitting
also measured the spectrum of the lamp near the end efficiency measurements, the xenon lamp with water
of the tube life (1337 h), and near the start of the life filter, and sunlight (at midday on 28 May 2004 in
2004 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

10

Spectral photon flux (1018 m-2 s-1 nm-1)


9

8 Tube 2, 5.7 h of operation


Tube 1, 1250 h of operation
7 Tube 1, 1337 h of operation
6

0
200 300 400 500 600 700 800
Wavelength (nm)

Fig. 3. Spectral photon flux for the xenon lamp with the water filter, for two different tubes and for three different tube ages. In all cases,
the total irradiance is normalised to 1000 W m−2 . Measurements were performed using a monochromator with a 12 nm bandpass.

potentiostat reference electrode. A Utah Electronics Model 0152 po-


A tentiostat was used to control the voltages. The elec-
trodes were immersed in 5M KOH electrolyte. A glass
-V W-R
frit separated the counter electrode from the working
and reference electrodes. Gaseous oxygen was purged
W R C from the counter electrode compartment using a con-
VW-C
tinuous flow of bubbled nitrogen gas. The presence of
N2(g)
the glass frit did not make a significant difference to
the photocurrent. However, allowing oxygen to dissolve
in the electrolyte surrounding the counter electrode led
to an increased photocurrent at a given cell bias volt-
quartz window
working age (VW.C in Fig. 4), by up to a factor of two at low
hv Pt
SCE
reference counter
bias voltages. This is attributed to the reduction of oxy-
porous glass frit gen molecules rather than the generation of hydrogen
molecules at the counter electrode. Measurements per-
Fig. 4. Schematic of the photoelectrochemical cell used for xenon formed in an oxygen-rich electrolyte can thus lead to
lamp measurements. A simplified equivalent circuit of the potentio- erroneously large efficiency values.
stat is included. VW.C and VW.R denote voltmeters measuring the The cell was constructed of glass, with a quartz win-
voltages between the working and counter-electrodes, and the work- dow to allow the incident radiation to enter with minimal
ing and reference electrodes, respectively. A denotes an ammeter
attenuation. The working electrode was placed 10 mm
measuring the photocurrent.
from the window. Swapping the positions of the work-
ing and reference electrodes had a negligible effect on
Sydney, Australia). Details of these sources were given the photocurrent, showing that attenuation by the elec-
in Section 2.3. trolyte and the resistance of the electrolyte were both
The xenon lamp measurements were performed using insignificant.
a standard three-electrode photoelectrochemical cell, The sunlight measurements were also performed us-
shown in Fig. 4. The three electrodes were the work- ing a three-electrode cell. In this case, the electrodes
ing electrode (the oxidised titanium sample), a plat- were immersed in 5M KOH electrolyte in a shallow
inum wire counter electrode, and a saturated calomel tray. Sunlight was directly incident on the electrolyte,
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2005

and the working electrode was oriented so that it faced irradiance was measured to be 880 ± 90 W m−2 . This
the sun. The counter and reference electrodes were the agrees, within the uncertainty of measurement, with the
same as for the xenon lamp measurements, and a flow 960 W m−2 calculated in Section 2.3.2.
of nitrogen gas was again used to purge oxygen from Fig. 5 shows the measured photoconversion efficien-
the counter electrode. cies for different bias voltages. The maximum efficiency
The irradiance was measured using an Oriel 71751 was found to be 0.54% at a bias voltage of 0.6 V for
sapphire window thermopile. The thermopile was cali- xenon lamp illumination, and 0.28% at the same bias
brated against a standard thermopile by the Australian voltage for solar illumination. The experimental uncer-
National Measurement Institute, using a standard lamp tainty for the measured photoconversion efficiencies is
with filament temperature of approximately 2700 K as estimated to be ±20%. The random repeatability er-
the light source. The uncertainty in the calibration is ror, which is a consequence of variation in factors such
±2%; spatial variations in the incident light intensity as electrolyte purity, working electrode alignment, and
increased the uncertainty in irradiance measurement to time for the photocurrent to reach equilibrium, was ap-
10%. proximately 15%. The largest measurement error was
The cell efficiency  for the water splitting reaction the 10% uncertainty in the irradiance measurements by
was determined using the thermopile.
 = jP (VWS − VB )/ES , (3) 2.5. Measurement of IPCE
where jP is the photocurrent produced per unit ir-
radiated area, VWS = 1.229 V is the potential corre- The IPCE provides a measure of the efficiency of con-
sponding to the Gibbs free energy change per photon version of photons incident on the photoelectrochem-
required to split water, and VB is the bias voltage ical cell to photocurrent flowing between the working
applied between the working and counter electrodes and counter electrodes. An IPCE of 100% corresponds
(VW.C in Fig. 4). The irradiance ES was measured to the generation of one photoelectron for each incident
at a position outside the cell. The quartz window photon. Losses associated with the reflection of inci-
used in the xenon lamp measurements reflects 4% dent photons, their imperfect absorption by the semi-
of this irradiance. No correction is made for this conductor, and recombination of charge carriers within
loss, since a window is likely to be required in a the semiconductor before they reach the electrolyte, all
working water-splitting reactor. In the solar mea- contribute to the IPCE values falling below 100%.
surements, no window is used; however, the water The IPCE was measured by interposing a monochro-
is at an angle to the direct solar radiation equal to mator between the xenon lamp and the photoelectro-
the solar zenith angle of 55◦ . We calculate that 7% chemical cell. Similar results were obtained both with
of the direct solar radiation will be reflected by the and without the use of a water filter. The bandpass of
water. To ensure comparability with the lamp mea- the monochromator was 12 nm (FWHM), which is suf-
surements, the measured solar irradiance was reduced ficiently narrow to have insignificant effect on the mea-
by 3%. sured IPCE values. The irradiance within the pass band
Some workers [5,7,8,11] have used the voltage be- was measured using the thermopile, averaging the mea-
tween the working and reference electrodes VW.R (rela- surement over a 20 s period. The IPCE was then calcu-
tive to the open-circuit value of this voltage) in place of lated as a function of wavelength using
the bias voltage VB . This is often the only voltage that
IPCE() = jP ()/[eI ()], (4)
is reported. However, the bias voltage is applied across
the working and counter electrodes, and the photocur- where I () is the incident photon flux passed by the
rent flows between these electrodes, so the electrical monochromator when set to wavelength , and e is the
power loss that has to be subtracted in calculating the electronic charge. The uncertainty of measurement was
efficiency of the cell is the product of the photocurrent calculated by combining the standard error on the time
and the voltage applied across the working and counter average of the irradiance with the errors noted for effi-
electrodes. Typically VW.R < VB , so using VW.R instead ciency measurements in Section 2.4.
of VB leads to an overestimate of the cell efficiency. Fig. 6 shows IPCE() for the TiO2 photoelectrode at
The photocurrent was measured for a range of ap- an applied bias voltage of 0.60 V, at which the maxi-
plied bias voltages, and the photoconversion efficiency mum photoconversion efficiency is obtained. The IPCE
calculated using Eq. (3). The irradiance for the xenon reaches a maximum of just under 60% for a wave-
lamp illumination was 820 ± 80 W m−2 . The solar length of 320 nm. The decrease at shorter wavelengths
2006 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

0.7
Xenon lamp with water filter
Sunlight, 28 May 2004, midday
0.6

Photoconversion efficiency (%)


0.5

0.4

0.3

0.2

0.1

0.0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2
Bias voltage (V)

Fig. 5. Photoconversion efficiency for the cell with TiO2 electrode as a function of bias voltage VB . Results are given for illumination by the
xenon lamp with water filter, and the sun.

70

60

50

40
IPCE (%)

30

20

10

0
300 320 340 360 380 400 420 440
Wavelength (nm)

Fig. 6. Measured incident photon conversion efficiency, expressed as a percentage, as a function of wavelength for the TiO2 photoelectrode.

is associated with an increase in reflection from the pho- assuming that the photocurrent is not dependent on the
toelectrode (see Section 3.3). At longer wavelengths, irradiance. The photoconversion efficiency can then be
weaker absorption of photons and the recombination of determined using
charge carriers lead to a decreased IPCE. These factors  ∞
are considered in detail in Section 3.4.  = e(VWS − VB ) IPCE()I () d/ES . (5)
Using the IPCE measurements, it is possible to cal- 0
culate the photocurrent expected to be generated for The cell photoconversion efficiency for rutile TiO2 cal-
illumination by light with any spectral dependency, culated for different spectra is given in Table 1.
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2007

Table 1
Photoconversion efficiencies for cells with rutile TiO2 photoanodes for different incident light sources

Spectrum Photoconversion efficiency

From IPCE (Eq. (5)) (%) Directly measured (Eq. (3)) (%)

AM1.5 global sunlight 0.25 ± 0.05


Sunlight, midday, 28 May 2004, Sydney 0.21 ± 0.04 0.28 ± 0.06
Xenon lamp without water filter (1250 h) 0.32 ± 0.07
Xenon lamp with water filter (1250 h) 0.51 ± 0.10 0.54 ± 0.11
Xenon lamp with water filter (5.7 h) 0.66 ± 0.15 0.66 ± 0.12
Xenon lamp with water filter (1337 h) 0.30 ± 0.06

Comparison of directly measured values with efficiencies calculated from the measured IPCEs integrated over the different source spectra.

The values given in Table 1 for sunlight and the xenon hydrogen molecule is produced for each pair of such
lamp with the water filter can be compared to the pho- electrons. The absorbed photon flux is then
toconversion efficiencies measured directly with those  g
sources. These efficiencies, given in Section 2.4, are re- JS = I () d (6)
peated in Table 1 and are equal to the values derived 0
from the IPCE measurements within the experimental
where g =hc/Ug is the band-gap wavelength. An upper
uncertainty.
bound to the photoconversion efficiency is given by
The age of the xenon lamp tube, and the presence of
the water filter, can strongly influence the measured cell ⎧
⎨ JS G (1 − loss )
0
photoconversion efficiency for rutile TiO2 . In particular, if Ug G0 + Uloss ,
C = ES (7)
the efficiency determined using the lamp tube near the ⎩
0 if Ug < G0 + Uloss ,
end of its lifetime (1337 h of operation) is much less
than determined using a new tube or a tube that has where G0 =eV WS =1.229 eV is the Gibbs free energy
been operated for 1250 h. The presence of a water filter change per electron for the water-splitting reaction. The
increases the measured efficiency by around 50%. factor loss represents the radiative quantum yield; i.e.,
the ratio of re-radiated photons to absorbed photons.
Hence 1−loss gives the proportion of absorbed photons
3. Calculated photoconversion efficiency converted into conduction band electrons. The reradia-
tion of photons is due to blackbody radiation from the
There are fundamental thermodynamic limits to excited state. Bolton et al. [26] have calculated that the
the photoconversion efficiency, which depend on the value of loss that corresponds to a maximum value of
band-gap of the semiconductor being used, and on the the efficiency is given by
spectrum of the incident radiation. In practice, these
limiting efficiencies cannot be achieved, due to losses loss P ≈ 1/ ln(JS /JBB ), (8)
such as the incomplete absorption of the incident ra-
where JBB is the blackbody photon flux at wavelengths
diation in the semiconductor, reflection of the incident
below the band-gap wavelength. loss P is small, less
radiation from the surface of the semiconductor and
than 0.02 for the range of band-gap wavelengths of in-
other surfaces, and recombination of conduction band
terest, and is practically independent of the spectrum of
electrons before reaching the electrolyte. In this sec-
the incident radiation.
tion we consider first the thermodynamic limits to
The term Uloss is the energy lost per photon. This is
efficiencies, and then the relevant loss factors.
a significant loss term, and provides a lower limit to the
band-gap energy of a semiconductor that can be used
3.1. Thermodynamic limits to efficiency for water splitting. There are two major contributions
to Uloss . The first is thermodynamic, and can be shown
We consider here an idealised semiconductor, in [26] to be equal to T Smix , where T is the temperature
which all photons of energy greater than the band- and Smix is the increase in entropy of mixing associ-
gap energy Ug are absorbed. The resulting conduction ated with production of conduction band electrons by
band electrons are transferred to the electrolyte, and a photon absorption. This contribution is at least 0.4 eV.
2008 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

35

30

Xenon lamp, without water filter

Maximum efficiency (%)


25 Xenon lamp, with water filter
AM1.5 global sunlight
20

15

10

Maximum
5 band-gap
wavelength
0
200 250 300 350 400 450 500 550 600 650 700 750
Band-gap wavelength (nm)

Fig. 7. The maximum photoconversion efficiency possible as a function of semiconductor band-gap wavelength. Results are given for AM1.5
global solar illumination, and illumination by the xenon arc lamp with and without the water filter. The maximum band-gap wavelength of
610 nm is indicated; materials with greater band-gap wavelengths cannot be used to split water.

The second contribution is a kinetic loss due to the over- maximum efficiency is greater for incident radiation
potentials for oxygen and hydrogen production. This from the xenon lamp, both with and without a water
has been estimated [27] to be at least 0.4 eV. There- filter, than for AM1.5 solar radiation. At band-gap
fore, we estimate Uloss 0.8 eV, which is in accordance wavelengths above 410 nm, the maximum efficiency
with Bolton et al.’s assessment [28]. This increases the obtainable with the xenon lamp without a water filter
minimum band gap energy from G0 = 1.23 eV to falls below that of AM1.5 solar radiation, while the
G0 +Uloss 2.03 eV, corresponding to a limiting band- maximum efficiency for the xenon lamp with a water
gap wavelength of 610 nm. filter remains larger for all band-gap wavelengths.
Note that there are two other loss factors implicit in Table 2 gives the maximum photoconversion efficien-
Eq. (7). First, although only photons with energy greater cies obtainable for a range of semiconductors of inter-
than Ug can be utilised to excite electrons from the va- est in water splitting, for the different radiation sources.
lence band to the conduction band, the fraction of the It can be seen that the maximum photoconversion effi-
photon energy greater than Ug is lost through rapid ther- ciencies for all materials are overestimated when using
mal equilibration with the semiconductor lattice. Sec- a xenon lamp fitted with a water filter. The overesti-
ond, even though Ug has to be significantly greater than mate is particularly severe for wide band gap materials
the energy G0 that is ultimately stored as chemical such as SrTiO3 and TiO2 . Use of a xenon lamp without
energy, the difference between Ug and G0 is lost. a water filter leads to overestimates of the efficiencies
Fig. 7 shows the maximum efficiency for water split- of wide band-gap materials, and underestimates of the
ting, calculated using Eq. (7), with loss =loss P . A cut- efficiencies of narrow band-gap materials.
off band-gap wavelength corresponding to Uloss =0.8 eV The efficiencies under xenon lamp illumination and
is indicated on the graph. The efficiency increases with the efficiency ratios will vary depending on the tube
band-gap wavelength, since more photons with energies age and other characteristics of the xenon lamp. Using
greater than the band-gap energy are available as the the spectra shown in Fig. 3 we calculate that the max-
band-gap wavelength increases. The relative efficien- imum efficiency obtainable for cells using rutile TiO2
cies of different radiation sources at a given band-gap electrodes is 4.8% for the new tube (5.7 h operation),
wavelength depend on the proportion of the photon flux and 2.3% for the tube at the end of its lifetime (1337 h
at wavelengths shorter than the band-gap wavelength. operation). These compare with 3.8% for the tube that
For band-gap wavelengths below about 410 nm, the has been used for 1250 h. Hence, the ratio of the AM1.5
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2009

Table 2
Maximum possible photoconversion efficiencies for semiconductors of different band gaps, for AM1.5 global solar illumination, and illumination
by a xenon arc lamp with and without a water filter fitted

Material Band-gap Band-gap Maximum efficiency (%) Ratio of efficiencies

energy (eV) wave-length (nm) AM1.5 Xe lamp Xe lamp AM1.5/Xe lamp AM1.5/Xe lamp
without filter with filter without filter with filter

SrTiO3 3.70 335 0.22 0.69 1.0 0.32 0.22


Anatase TiO2 3.20 388 1.3 1.7 2.6 0.77 0.49
Rutile TiO2 3.00 413 2.2 2.3 3.8 0.98 0.60
WO3 2.70 459 4.8 3.7 6.2 1.32 0.78
CdS 2.40 517 9.1 6.0 10.4 1.52 0.87
TaON 2.38 520 9.3 6.1 10.7 1.53 0.88
C-doped TiO2 [8] 2.32 535 10.5 6.7 11.7 1.57 0.90
Fe2 O3 2.20 564 12.9 7.9 14.0 1.63 0.92
Ta3 N5 2.07 600 15.9 9.6 16.9 1.67 0.94
Hypothetical ideal material 2.03 610 16.8 10.0 17.7 1.68 0.95

efficiency to that for the xenon lamp with water filter 6.9 × 107 m−1 at 300 nm, but falls by almost four or-
can vary from 0.95 to 0.45 for rutile TiO2 , depending ders of magnitude to 9.1 × 103 m−1 at the band-gap
on the tube age. For most of the tube lifetime, the ratio wavelength of 413 nm.
is between 0.45 and 0.60. Fig. 8 shows the absorbed spectral photon flux
()I () for different thicknesses of rutile, calculated
3.2. Effect of imperfect absorption on efficiency using Eq. (9), for AM1.5 global solar radiation. It can
be seen that a thickness of about 1 mm is required
The results presented in Section 3.1 were calculated for essentially all photons with wavelengths below the
assuming the semiconductor absorbs all radiation at band-gap wavelength to be absorbed. Fig. 9 shows the
wavelengths below the band-gap wavelength; i.e., it maximum photoconversion efficiency, calculated using
was assumed that the absorbed photon flux is given by Eq. (7), for rutile semiconductors of different thick-
Eq. (6). The absorption coefficients of real semiconduc- nesses, taking into account the imperfect absorption.
tors in fact decrease as the wavelength approaches the Results are given for AM1.5 global solar radiation and
band-gap wavelength from below. To take this into ac- for the xenon lamp with and without a water filter. The
count, the absorbed photon flux should be calculated band-gap energy is assumed to be 3.0 eV (correspond-
using ing to wavelength 413 nm), for which C = 2.25% for
AM1.5 sunlight, if all radiation at wavelengths below
 g the band-gap wavelength is absorbed. Fig. 9, shows
JS = ()I () d, (9) that this maximum efficiency is only obtained for thick-
0
nesses approaching 1 mm. For a thickness of 1 m,
where the efficiency is 1.2%, and for a thickness of 0.1 m,
the efficiency is only 0.65%. For xenon lamp illumi-
() = 1 − exp[−k()h] (10) nation, particularly with the water filter, the efficiency
is greater, owing to the greater proportion of the more
is the fraction of incident radiation absorbed in the strongly absorbed UV radiation.
wavelength band from  to d, k() is the absorption The calculations presented here do not take into ac-
coefficient and h is the thickness of the semiconductor. count reflection from the rear surface of the TiO2 layer.
It is assumed that absorption only takes place at wave- The reflection coefficient from this surface is calcu-
lengths shorter than the band-gap wavelength g of the lated to be about 30% for the titanium substrate used in
semiconductor. the experiments. It is therefore likely that the calcula-
In this section, we investigate the effect of tak- tions presented here slightly underestimate the absorbed
ing into account real absorption coefficients, using power, particularly under conditions when a substantial
the data of Eagles [19] for the absorption coefficient proportion of the incident radiation is transmitted by the
of single-crystal rutile. The absorption coefficient is TiO2 layer.
2010 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

3.0
1 nm
10 nm

Spectral photon flux / 1018 (m-2 nm-1 s-1)


2.5 100 nm
1 µm
10 µm
2.0 100 µm
1 mm
AM1.5 global sunlight
1.5

1.0

0.5

0.0
300 320 340 360 380 400 420
Wavelength (nm)

Fig. 8. Wavelength dependence of the absorbed spectral photon flux for different thicknesses of rutile, for AM1.5 global solar illumination.

4.0
Xe lamp without water filter
3.5 Xe lamp with water filter
AM1.5 global sunlight
3.0

2.5
Efficiency (%)

2.0

1.5

1.0

0.5

0.0
10-9 10-8 10-7 10-6 10-5 10-4 10-3
Rutile thickness (m)

Fig. 9. Dependence of maximum photoconversion efficiency on thickness of rutile layer, taking into account absorption of photons in the
rutile. Results are given for AM1.5 global solar illumination, and illumination by the xenon arc lamp with and without the water filter.

3.3. Effect of reflection on efficiency For collimated radiation passing from medium 1 to
medium 2 at normal incidence to the boundary between
A proportion of the incident light is reflected at any the media, the reflection coefficient is given by
interface at which the refractive index changes. In a typ-
ical water-splitting reactor, a semiconductor electrode (n2 − n1 )2 + (2 − 1 )2
is immersed in an aqueous solution of a salt, an acid R12 = , (11)
(n2 + n1 )2 + (2 + 1 )2
or a base, and light enters through a quartz window.
Therefore reflection occurs at the interface between the where Ni = ni + ii is the complex refractive indices of
air and the quartz, the quartz and the solution, and the medium i. Here ni is the real part of the refractive index
solution and the semiconductor. (usually referred to as simply the refractive index) and
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2011

0.40

0.35
Air-quartz
Quartz-water
0.30
Water-TiO2
Reflection coefficient 0.25 Total

0.20

0.15

0.10

0.05

0.00
250 300 350 400 450 500 550 600
Wavelength (nm)

Fig. 10. Reflection coefficients at the interfaces between air and quartz glass, quartz glass and water, and water and rutile TiO2 . The coefficients
are for specular reflection at normal incidence.

i is the extinction coefficient, related to the absorption where R() is the total reflection coefficient (i.e., the
coefficient k by k = 4i /. sum of the individual reflection coefficients). Fig. 11
We consider here as an example the reflection losses shows the influence of reflection on the photoconver-
for a TiO2 electrode immersed in water, with a quartz sion efficiency for different thicknesses of TiO2 . Reflec-
window between the water and the air. This was the tion leads to a decrease in efficiency of between 24%
arrangement used for the xenon lamp measurements. It and 33%, depending on the thickness of the TiO2 . The
was noted in Section 2.4 that the solar measurements decrease is essentially independent of the light source.
were performed in a vessel with no quartz window; Reflection reduces the maximum efficiency for AM1.5
however, it is likely that reactors will use such a window. sunlight to 1.7% for a thick rutile layer, and to 0.45%
Refractive index data for water and quartz were taken for a 0.1 m layer.
from the SOPRA optical constant database [29], and for There are several sources of uncertainty in the reflec-
rutile TiO2 as noted in Section 2.2. The TiO2 absorption tion coefficient calculations. First, the refractive index
coefficient was assumed to be constant at wavelengths of water has been used instead of the refractive index of
below 300 nm, as suggested by the data of Eagles [19]. a basic solution. However, this makes only a small dif-
The relevant reflection coefficients, calculated using ference to the calculated reflection coefficients. The re-
Eq. (11), are shown in Fig. 10. Reflection losses at fractive index of water at 589 nm is 1.333, while that of
the interface between the quartz glass and water are 5M KOH is 1.376 [30]. The total reflection coefficient
very low, and those between the air and quartz glass at 589 nm is reduced from 15.9% to 14.8% when the re-
are about 4%. The reflection losses at the TiO2 surface fractive index of water is replaced by that of 5M KOH.
are substantial, reaching more than 30% at short wave- Second, the reflection coefficients have been cal-
lengths. It is interesting to note that while large absorp- culated for normal incidence. Reflection coefficients,
tion coefficients are required to minimise absorption and therefore losses, increase as the angle of incidence
losses, they have the undesirable effect of increasing (which is zero at normal incidence) increases. For
reflection losses. global solar irradiation, a proportion of the radiation
The influence of reflection on the efficiency of con- is scattered from the atmosphere, and therefore has a
version of solar energy to stored chemical energy can non-zero angle of incidence even when the electrode is
be calculated by modifying the absorbed photon flux facing the sun. Hence, reflection losses will be larger
JS : than those presented in Fig. 11. We estimated the
 g influence of this by calculating diffuse reflection coef-
ficients (averaged over all angles of incidence) for the
JS = [1 − R()]()I () d, (12)
0 air–quartz, quartz–water and water–rutile interfaces,
2012 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

Efficiency with reflection / efficiency without reflection


4.0 Xe lamp without filter 1.0
Xe lamp with filter
3.5 AM1.5 global solar
0.8
3.0

2.5
0.6
Efficiency (%)

2.0

0.4
1.5

1.0
0.2
0.5

0.0 0.0
10-9 10-8 10-7 10-6 10-5 10-4 10-3
Rutile thickness (m)

Fig. 11. Dependence of photoconversion efficiency on thickness of rutile layer, taking into account reflection from the quartz, water and rutile,
and absorption of photons in the rutile. Results are given for AM1.5 global solar illumination, and illumination by the xenon arc lamp with
and without the water filter. Also shown is the ratio of photoconversion efficiency achievable including and excluding reflection.

taking into account refraction of the rays at the first two fusion length of charge carriers in the TiO2 . We use
interfaces. The reflection at the first interface increases the simple Schottky-barrier model presented by Ghosh
significantly, resulting in an increase in the total reflec- and Maruska [32,33], in which the semiconductor layer
tion coefficient by a factor of about 30%. We applied of thickness h is assumed to behave like a conven-
these reflection coefficients to the diffuse component of tional Schottky barrier. There is a barrier region (carrier
the global AM1.5 spectrum, and the normal incidence depletion, or space charge, region) of width lb at the
reflection coefficients to the remainder of the spectrum. semiconductor–electrolyte interface. All carriers pro-
The photoconversion efficiency decreased by a factor duced in the barrier region, or that reach the barrier re-
of between 9% (for a 1 nm thick rutile layer) and 6% gion by diffusion, are assumed to reach the electrolyte.
(for a 1 mm thick rutile layer). Again, these changes Carriers produced outside the barrier region (i.e., a dis-
are relatively small. tance greater than lb from the semiconductor–electrolyte
Finally, the reflection coefficients have been calcu- interface) have to diffuse through the bulk of the semi-
lated assuming that the TiO2 layer has a smooth surface, conductor, with diffusion length L, to reach the barrier
while the surface is rough, which may reduce the to- region. A similar model was used by Butler [34].
tal reflectance, integrated over a hemisphere. Scanning The number of carriers produced per unit time within
electron microscope images show that the rms rough- the interval dx at a distance x from the surface is pro-
ness of the surface is much less than 100 nm. Calcula- portional to GI k exp(−kx) dx, where G is the quan-
tions using a physical optics model [31] indicate that tum conversion efficiency. Assuming that G is constant
the total reflectance from such a surface is at least 97% with respect to x and writing  = 1/L gives
of that from a smooth surface. 
k
IPCE = G [1 − exp(−kl b )] + exp(lb )
k+
3.4. Calculation of diffusion length of charge carriers 
from IPCE measurements × {exp[−(k + )]lb − exp[−(k + )]h} .
(13)
The measured incident photon conversion efficiency
for the TiO2 electrode, shown in Fig. 6 was used in A least-squares fit of the IPCE, calculated using (13),
Section 2.5 to calculate the photoconversion efficiency to the measured data shown in Fig. 6 was performed.
under illumination by different light sources. In this The three terms that were fitted were L, lb and G. The
section, we use the measured IPCE to estimate the dif- measured IPCEs were divided by (1−R) to compensate
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2013

measured
100
model, best fit
90 (L = 180 nm, lb = 8 nm, G = 0.96)
barrier contribution
80
bulk contribution
70 α(λ)

60
IPCE (%)

50

40

30

20

10

0
320 340 360 380 400 420
Wavelength (nm)

Fig. 12. Measured IPCE at 0.6 V bias voltage, modified to compensate for reflection losses, for the flame-oxidised TiO2 electrode, with the
best fit of the Schottky barrier diode model to the measurements. The contribution of the barrier region and the bulk region to the IPCE are
shown for the best fit. Also shown is , the percentage of the radiation absorbed in the TiO2 layer.

for reflection losses. The thickness h = 280 ± 40 nm The value of L = 180 ± 30 nm indicates that only
measured by reflectometry was used. The best-fit values photons absorbed in about the top 200 nm of the TiO2
of lb , L and G were found to be 8 ± 7 nm, 180 ± can produce charge carriers that are used in electrolysis.
30 nm and 0.96±0.05, respectively, for h=280 nm. The This suggests that, for the present sample, there is little
uncertainties reflect the values for which the fit becomes benefit in having a layer of thickness much greater than
noticeably less accurate. The values are only weakly 200 nm, and that diffusion of carriers through the bulk
dependent on h; for example, varying h by ±40 nm is the main contribution to photocurrent generation.
altered L by only ±10 nm.
The calculated dependence of IPCE on wavelength
is shown in Fig. 12 together with the components due 4. Discussion
to photons absorbed in the barrier region (x lb ) and
the bulk region (lb < x L). These components corre- 4.1. Comparison of measured and calculated
spond respectively to the first and second terms on the efficiencies
right-hand side of Eq. (13). The agreement between
the calculation and the measured data is reasonable for The photoconversion efficiencies calculated for dif-
wavelengths up to 380 nm; it is not clear whether the ferent radiation sources are compared with measured
discrepancy at longer wavelengths is a consequence of efficiencies in Table 3. For each source, six calculated
deviations of the absorption coefficient from the single- efficiency figures are given. The first is the theoretical
crystal rutile data that was used, or deficiencies in the maximum value, calculated using Eq. (7). The second
simple Schottky barrier model. Also shown in Fig. 12 is takes into account losses due to imperfect absorption,
(), calculated using Eq. (10) with h = 280 nm, which and the third due to reflection as well as imperfect ab-
corresponds to the IPCE assuming that all photons ab- sorption and charge transfer. The influence of the bias
sorbed within the thickness of the TiO2 layer partici- voltage of 0.6 V that is applied in the measurements is
pate in electrochemical reactions. The measured IPCE taken into account in the fourth figure using
is significantly smaller than this curve for wavelengths
above 330 nm, indicating that recombination of charge  = (VWS − VB )/1.23 V, (14)
carriers is significant. For shorter wavelengths, where
absorption occurs close to the surface of the semicon- where  and  are, respectively, the efficiencies cal-
ductor, the quantum yield is consistent with the radiative culated excluding and including the effect of the bias
limit of (1 − loss P ) ≈ 0.98 calculated in Section 3.1. voltage.
2014 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

Table 3
Comparison of measured photoconversion efficiencies for rutile TiO2 with those calculated for various effective rutile thicknesses h , reflection
coefficients R and bias voltages VB

Efficiency AM1.5 global Sunlight, 28 May 2004, Xenon lamp with Xenon lamp
sunlight midday, Sydney water filter without water filter

Calculated
h = ∞ 2.25% 1.94% 3.75% 2.29%
R=0
VB = 0

h = 280 ± 40 nm 0.92 ± 0.04% 0.76 ± 0.04% 2.09 ± 0.06% 1.35 ± 0.04%


R=0
VB = 0

h = 280 ± 40 nm 0.66 ± 0.03% 0.54 ± 0.03% 1.47 ± 0.05% 0.94 ± 0.03%


R from Fig. 10
VB = 0

h = 280 ± 40 nm 0.34 ± 0.02% 0.28 ± 0.02% 0.75 ± 0.02% 0.48 ± 0.01%


R from Fig. 10
VB = 0.6 V

h = 180 ± 30 nm 0.57 ± 0.03% 0.47 ± 0.03% 1.34 ± 0.06% 0.86 ± 0.03%


R from Fig. 10
VB = 0

h = 180 ± 30 nm 0.29 ± 0.02% 0.24 ± 0.02% 0.69 ± 0.03% 0.44 ± 0.02%


R from Fig. 10
VB = 0.6 V

Measured
From IPCE 0.25 ± 0.05% 0.21 ± 0.04% 0.51 ± 0.10% 0.32 ± 0.07%
Direct 0.28 ± 0.06% 0.54 ± 0.11%

The first four figures are calculated assuming that the measured and calculated efficiencies. Similar cal-
the quantum efficiency reaches its maximum possible culations for different semiconductors would provide a
value, so that all charge carriers produced in the full useful guide to the achievable efficiencies.
thickness of the TiO2 layer participate in water-splitting Of particular interest is the large difference between
reactions. The fifth and sixth figures take into account the theoretical maximum efficiency, and that achievable
charge transfer between the semiconductor and the elec- once loss mechanisms are taken into account. The losses
trolyte, using an effective TiO2 layer thickness equal to are of the order of a factor of 7 to 8 for sunlight. This
the diffusion length of 180 ± 30 nm derived in Section is a major concern, given that the theoretical maximum
3.4. The sixth figure, which takes into account the bias efficiency for a semiconductor with an ideal band-gap
voltage as well, is the best estimate of efficiency from energy of 2.0 eV is only 17%. Clearly to obtain the
our calculations. It is compared in Table 3 to the effi- target efficiency of 10% will require substantial effort
ciencies measured both directly and derived from the to minimise losses. If possible, the flat-band potential
measured IPCE. It can be seen that the calculated effi- of the conduction band should be more negative than
ciency for sunlight agrees well with the measurements, the hydrogen redox potential to remove the need for a
while that for the xenon lamp is slightly higher. bias voltage. A thick nanostructured or mesostructured
There are many uncertainties in the calculated effi- semiconductor photoelectrode could be used to ensure
ciencies, such as the influence of roughness of the TiO2 all the incident light is absorbed, while allowing the
surface on reflection coefficients, the use of absorption electrolyte to penetrate the semiconductor to minimise
and reflection coefficients for single crystal rutile, the recombination losses [35].
neglect of absorption of radiation reflected from the
TiO2 –Ti interface, and the use of a simple model to 4.2. Influence of the light source on efficiency
calculated the effective thickness of the film. These un- Our results show that the spectrum of the light source
certainties account for the small discrepancies between has a large impact on the measured photoconversion
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2015

efficiency. We have investigated the most widely-used underestimated the solar irradiance as 290 W m−2 , ob-
light source, an ozone-free xenon lamp. The mea- taining an efficiency of 0.4% for TiO2 produced by
sured efficiency is overestimated, relative to that for flame oxidation; using a more realistic irradiance would
the AM1.5 global solar spectrum, for wide band- reduce the efficiency to well below 0.2%.
gap semiconductors and when a water filter is fitted. The most reliable published solar measurement is
Other commonly-used lamps, such as mercury lamps, probably that of Ghosh and Maruska [32] who obtained
mercury–xenon lamps and non ozone-free xenon lamps an efficiency of 0.6% for aluminium-doped TiO2 un-
have spectra more heavily weighted to the ultraviolet der solar illumination for an irradiance of 1.05 kW m−2 .
than ozone-free xenon lamps, and will lead to more The efficiency for undoped TiO2 was about 35% of the
severe overestimates of efficiency. Al-doped TiO2 , and is hence about 0.21%, which is
Even when the spectrum of the light source is known, consistent with our result.
it is not possible to derive a universal conversion factor
to convert measured photoconversion efficiencies to ef- 4.3.2. Measurements under artificial illumination
ficiencies under AM1.5 solar illumination. The conver- It was noted in Section 1 that many workers have
sion factor would depend on the band gap, and on the quoted efficiencies obtained under artificial illumina-
wavelength dependence of the absorption and reflection tion, and that some of these efficiencies are close to or
characteristics of the semiconductor, and hence would higher than the theoretical thermodynamic limits for the
vary from material to material. A further difficulty is semiconductors used. Here we examine some specific
that the spectrum of a given type of lamp depends on cases in the light of our results.
its age. We have shown that the conversion factor can Khan et al. [8] reported obtaining an efficiency of
vary by a factor of around two, depending on the age 8.35% with carbon-doped TiO2 with an extended band-
of the tube. gap wavelength of 535 nm. Table 2 shows that the max-
Our results indicate that measurement of efficiency imum efficiency under AM1.5 global solar illumina-
under AM1.5 solar illumination has to be performed ei- tion is 10.5% for this band-gap wavelength. The stated
ther directly, or by measuring the IPCE as a function of bias voltage was 0.3 V; taking this into account using
wavelength and integrating over the AM1.5 spectrum. Eq. (14) gives a maximum efficiency of 8.0%. A simi-
There are significant uncertainties for solar measure- lar maximum efficiency figure of 8.1% was calculated
ment, since the spectrum can vary depending on the by Hägglund et al. [39]. Losses due to imperfect ab-
zenith angle of the sun, and atmospheric conditions, par- sorption, reflection and recombination will almost cer-
ticularly the ozone column, and ground conditions. We tainly reduce the efficiency to below 2%. It should also
have found, for example, differences of 15% to 20% in be noted that Khan et al. measured the bias voltage
the efficiencies calculated for AM1.5 global spectrum between the working electrode and the reference elec-
and for the solar spectrum relevant to our measurement trode, and defined the cell bias voltage as the differ-
performed under solar illumination. Probably the most ence between this in the open and closed circuit state.
reliable means of relating measured efficiencies to the The actual bias voltage was likely to be closer to 0.5 V
AM1.5 global solar spectrum, or any other standard than the stated 0.3 V. This will decrease the efficiency
spectrum, is to measure the IPCE and integrate over the by a further factor of around 20%. Additionally, no pre-
chosen spectrum using Eq. (5) cautions were taken to eliminate oxygen from the elec-
trolyte surrounding the counter electrode, which may
4.3. Relevance to other work cause a further overestimate of efficiency.
Khan et al. used a xenon arc lamp with an infrared
4.3.1. Measurements under solar illumination filter as their light source. Table 2 indicates that this will
Few previous publications have compared photocon- contribute strongly to the discrepancy between the re-
version efficiencies obtained under solar and artificial il- ported efficiency of 8.35% and the maximum efficien-
lumination. Houlihan et al. [36] presented data for TiO2 cies estimated here. It is also possible that the spectral
and doped TiO2 electrodes, finding larger efficiencies response of the photometer used by Khan et al. to mea-
under sunlight than under xenon lamp illumination (for sure the irradiance was weighted to short wavelengths,
example, 0.70% under sunlight and 0.26% under xenon which would lead to an underestimate of the incident
lamp for a TiO2 electrode). However, the solar irradi- power and therefore an overestimate of the photocon-
ance was assumed to be 400 W m−2 at around midday version efficiency. We note that Khan et al. [8] obtained
in June in Pennsylvania, USA [37]; this is too low by an efficiency of 1.08% for oven-oxidised TiO2 , with a
a factor of around 2.5. Fujishima et al. [38] similarly standard rutile band gap of 3.0 eV. This is substantially
2016 A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017

above the values we presented in Table 3 even for xenon sources, without taking into account the differences
arc lamp illumination. The photoconversion efficiencies between the spectra of the light source and the AM1.5
we have obtained for oven-oxidised TiO2 have gener- solar spectrum. Other problems are also evident in
ally been lower than those we have obtained for flame- some efficiency measurements, such as the use of the
oxidised TiO2 . voltage between the working and reference electrodes,
Other workers have also presented photoconversion rather than the working and counter electrodes, as the
efficiencies above those possible under AM1.5 global bias voltage, and incorrect measurement of the total
solar illumination. For example, the 2.5% reported by irradiance.
Tang et al. [10] for an anatase photoelectrode under We have calculated the thermodynamically achiev-
xenon lamp illumination is clearly greater than the 1.3% able photoconversion efficiency for semiconductors of
possible for AM1.5 illumination (and the 1.7% possible different band gaps, and investigated the absorption,
for a xenon lamp without a water filter). The discrepancy reflection and recombination losses for rutile. These
here is due to an underestimate of the total irradiance losses, and the loss due to the requirement for a bias
[40]. Mishra et al. [9] reported efficiencies up to 3% voltage, are found to account for the discrepancy be-
for TiO2 photoelectrodes using a mercury–xenon lamp tween the theoretically achievable efficiency and the
source; again this is much higher than is possible for measured efficiency. The losses lead to a major reduc-
AM1.5 illumination. The 1.9% efficiency obtained by tion in the efficiency, and will have to be minimised if
Radecka et al. [11] using a xenon lamp for a mixed photoconversion efficiencies approaching 10% are to
anatase-rutile photoelectrode is greater than is possible be obtained.
for anatase or rutile under AM1.5 illumination when the
stated 0.2 V bias voltage is taken into account. An earlier
paper by Khan and Akikusa [5] quoted an efficiency of Acknowledgments
2.0% for flame-oxidised titanium with a bias voltage of
0.7 V; the maximum achievable under AM1.5 sunlight We thank Mr. Paul Gwan of CSIRO Industrial Physics
is 0.9% in this case. for assistance with preparing the rutile electrode, Ms.
A number of workers have presented IPCE measure- Christina Li of CSIRO Minerals for assistance with the
ments, but have not taken the next step of integrating IPCE measurements, and Mr. Errol Atkinson and Dr.
these data over the AM1.5 solar spectrum to obtain an Frank Wilkinson of the Australian National Measure-
efficiency that is easily comparable with the results of ment Institute for calibrating the xenon lamp and the
others. thermopile.

References
5. Conclusions
[1] Fujishima A, Honda K. Electrochemical photolysis of water at
We have demonstrated that the use of a xenon lamp a semiconductor electrode. Nature 1972;238:37–8.
[2] Bard AJ, Fox MA. Artificial photosynthesis: solar splitting of
as the light source can lead to significant overestimates water to hydrogen and oxygen. Acc Chem Res 1995;28:141–5.
of the photoconversion efficiencies, compared to effi- [3] Nozik AJ. Photoelectrolysis of water using semiconducting
ciencies under the standard AM1.5 solar illumination. TiO2 crystals. Nature 1975;257:383–6.
The overestimate is particularly large when a water fil- [4] Giordano N, Antonucci V, Cavallaro S, Lembi R, Bart JCJ.
Photoassisted decomposition of water over modified rutile
ter is used on the xenon lamp, and for wide-band gap
electrodes. Int J Hydrogen Energy 1982;7:867–72.
semiconductors such as TiO2 . We conclude that reliable [5] Khan SUM, Akikusa J. Stability and photoresponse of
estimation of efficiency under standard solar conditions nanocrystalline n-TiO2 and n-TiO2 /Mn2 O3 thin film electrodes
is best performed by measuring the IPCE at the cell bias during water splitting reactions. J Electrochem Soc 1998;145:
voltage of maximum efficiency as a function of wave- 89–93.
[6] Akikusa J. Investigation of thin film n-TiO2 , n-Fe2 O3 and single
length, and integrating over the AM1.5 solar spectrum. crystal p-Si, p-SiC semiconductors for photosplitting of water,
Measurements under sunlight should also provide rea- PhD thesis, Duquesne University, Pittsburgh, USA, 1996.
sonable estimates of efficiency, provided that the solar [7] Akikusa J, Khan SUM. Photoresponse and AC impedance
zenith angle is close to the 48.2◦ used for the AM1.5 characterization of n-TiO2 films during hydrogen and oxygen
spectrum. evolution reactions in an electrochemical cell. Int J Hydrogen
Energy 1997;22:875–82.
Many of the photoconversion efficiencies presented [8] Khan SUM, Al-Shahry M, Ingler Jr WB. Efficient
in the literature are of limited usefulness, since they photochemical water splitting by a chemically modified n-TiO2 .
have been measured with xenon lamps and other light Science 2002;297:2243–5.
A.B. Murphy et al. / International Journal of Hydrogen Energy 31 (2006) 1999 – 2017 2017

[9] Mishra PR, Shukla PK, Singh AK, Srivastava ON. Investigation [25] Oriel Instruments. The book of photon tools. CT: Stratford,
and optimization of nanostructured TiO2 photoelectrode in 1999.
regard to hydrogen production through photoelectrochemical [26] Bolton JR, Haught AF, Ross RT. Photochemical energy storage:
process. Int J Hydrogen Energy 2003;28:1089–94. an analysis of limits. In: Connolly JS, editor. Photochemical
[10] Tang J, Wu Y, McFarland EW, Stuckey GD. Synthesis and conversion and storage of solar energy (Proceedings of the
photocatalytic properties of highly crystalline and ordered third international conference, Boulder, CO, 1980). New York:
mesoporous TiO2 films. Chem Commun 2004;2004:1670–1. Academic Press; 1981. p. 297–339.
[11] Radecka M, Wierzbicka M, Komornicki S, Rekas M. Influence [27] Weber MF, Dignam MJ. Efficiency of splitting water
of Cr on photoelectrochemical properties of TiO2 thin films. with semiconducting photoelectrodes. J Electrochem Soc
Physica B 2004;348:160–8. 1984;131:1258–65.
[12] Licht S. Solar water splitting to generate hydrogen fuel: [28] Bolton JR, Strickler SJ, Connolly JS. Limiting and realizable
photothermal electrochemical analysis. J Phys Chem B efficiencies of solar photolysis of water. Nature 1985;316:
2003;107:4253–60. 495–500.
[13] Licht S. Solar water splitting to generate hydrogen fuel—a [29] SOPRA Optical Constant Database, http://www.sopra-sa.com/
photothermal electrochemical analysis. Int J Hydrogen Energy more/database.asp, accessed 27 May 2004.
2005;30:459–70. [30] Lide DR, editor. CRC handbook of chemistry and physics, 84th
[14] Barnes PRF, Randeniya LK, Murphy AB, Gwan PB, Plumb IC, ed. Boca Raton, FL: CRC Press, 2003. p. 8.59–85.
Glasscock JA. et al. TiO2 photoelectrodes for water splitting: [31] He XD, Torrance KE, Sillion FX, Greenberg DP. A
carbon doping by flame pyrolysis?. Dev Chem Eng Mineral comprehensive physical model for light reflection. Comput
Process 2006;14:51–70. Graph 1991;25:175–86.
[15] Heavens OS. Optical properties of thin films. London: [32] Ghosh AK, Maruska HP. Photoelectrolysis of water in sunlight
Butterworths, 1955. p. 76–7. with sensitized semiconductor electrodes. J Electrochem Soc
[16] Cardona M, Harbeke G. Optical properties and band structure 1977;124:1516–22.
of wurtzite-type crystals and rutile. Phys Rev 1965;137A: [33] Ghosh AK, Morel DL, Feng T, Shaw RF, Rowe Jr
1467–76. CA. Photovoltaic and rectification properties of Al/Mg
[17] DeVore JR. Refractive indices of rutile and sphalerite. J Opt phthalocyanine/Ag Schottky-barrier cells. J Appl Phys
Soc Am 1951;41:416–9. 1974;45:230–6.
[18] Ribarsky MW. Titanium dioxide (TiO2 ) (rutile). In: Palik ED, [34] Butler MA. Photoelectrolysis and physical properties of the
editor. Handbook of optical constants, Orlando, FL: Academic, semiconducting electrode WO3 . J Appl Phys 1977;48:1914–21.
1985. p. 795–800. [35] Santato C, Ulmann M, Augustynski J. Photoelectrochemical
[19] Eagles DM. Polar modes of lattice vibration and polaron properties of nanostructured tungsten trioxide films. J Phys
coupling constants in rutile (TiO2 ). J Phys Chem Solids Chem B 2001;105:936–40.
1964;25:1243–51. [36] Houlihan JF, Hamilton JR, Madacsi DP. Improved solar
[20] Mergel D. Modeling thin TiO2 films of various densities as an efficiencies for doped polycrystalline TiO2 photoanodes. Mat
effective optical medium. Thin Solid Films 2001;397:216–22. Res Bull 1979;14:915–20.
[21] ASTM. Standard tables for reference solar spectral irradiance at [37] Houlihan JF, Madacsi DP, Walsh EJ, Mulay LN. Improved solar
air mass 1.5: Direct normal and hemispherical for a 37◦ tilted energy conversion efficiencies for the photocatalytic production
surface. Standard G159-98. West Conshohocken, PA: American of hydrogen via TiO2 semiconductor electrodes. Mat Res Bull
Society for Testing and Materials, 1998. 1976;11:1191–8.
[22] Gueymard CA. Parameterized transmittance model for direct [38] Fujishima A, Kohayakawa K, Honda K. Hydrogen production
beam and circumsolar spectral irradiance. Solar Energy under sunlight with an electrochemical photocell. J Electrochem
2001;71:325–46. Soc 1975;122:1487–9.
[23] Gueymard C. SMARTS2, a simple model of the atmospheric [39] Hägglund C, Grätzel M, Kasemo B. Comment on “Efficient
radiative transfer of sunshine: Algorithms and performance photochemical water splitting by a chemically modified n-TiO2 ”
assessment. Professional Paper FSEC-PF-270-95. Cocoa, FL: (II). Science 2003;301:1673b.
Florida Solar Energy Center, 1995. [40] Tang J, Private communication, 2005.
[24] National Measurement Institute. Australian Government
Department of Industry, Tourism and Resources, P.O. Box 264,
Lindfield NSW 2070, Australia. http://measurement.gov.au

You might also like