Anda di halaman 1dari 12

ARTICLE IN PRESS

Solar Energy Materials & Solar Cells 91 (2007) 1326–1337


www.elsevier.com/locate/solmat

Band-gap determination from diffuse reflectance measurements of


semiconductor films, and application to photoelectrochemical
water-splitting
A.B. Murphya,b,
a
CSIRO Industrial Physics, P.O. Box 218, Lindfield, NSW 2070, Australia
b
CSIRO Energy Transformed National Research Flagship, PO Box 330, Newcastle, NSW 2300, Australia

Received 16 March 2007; received in revised form 4 May 2007; accepted 4 May 2007
Available online 20 June 2007

Abstract

Measurements of the diffuse reflectance of TiO2 semiconductor coatings, such as are used for water splitting, are analysed using the
Kubelka–Munk radiative transfer model. The widely used practice of determining the band gap of the coating directly from the diffuse
reflectance is found to be inaccurate, since the diffuse reflectance depends on parameters such as the thickness, refractive index and
surface roughness of the coating. However, it is shown that the absorption coefficient can be derived from the diffuse reflectance using an
inversion method; the band gap can then be obtained from the absorption coefficient. Finally, the diffuse reflectance of carbon-doped
TiO2 presented by Khan et al. [Science 297 (2002) 2243-2245] is analysed; it is found that while the band-gap wavelength is extended into
the visible region, it is overestimated. Moreover, light at visible wavelengths is only very weakly absorbed, and is expected to make only a
minor contribution to the water-splitting efficiency.
Crown Copyright r 2007 Published by Elsevier B.V. All rights reserved.

Keywords: Photocatalysis; Absorption coefficient; Tauc plot; Band gap; Solar hydrogen production

1. Introduction as large a proportion as possible of the solar spectrum is


utilized, it is important that the band gap of the
In photoelectrochemical water-splitting, hydrogen is semiconductor is as close as possible to the 2 eV required
produced in an electrochemical cell by the action of light for the charge carriers to have sufficient energy to split
on a photoelectrode, typically a metal-oxide semiconductor water [1–3].
thin film on a conducting substrate. The semiconductor Measurement of diffuse reflectance with a UV-visible
absorbs photons at wavelengths below its band-gap spectrophotometer is a standard technique in the determi-
wavelength, producing electron–hole pairs. These charge nation of the absorption properties of materials. In the case
carriers diffuse to the water and the conducting substrate, of semiconductors for water splitting, the properties that
driving the water-splitting reaction to produce hydrogen can potentially be estimated from the diffuse reflectance are
and oxygen. Since the diffusion length of the charge the band-gap energy (also referred to as the band gap) and
carriers is small (200 nm or less in titanium dioxide), it is the absorption coefficient. Determination of the band gap
important that the photons are absorbed near the surface; from the measurement of the diffuse reflectance of a
i.e., that the semiconductor has a large absorption powder sample is a standard technique [4,5]. The powder
coefficient for sub-band-gap wavelengths. Further, so that sample has to be sufficiently thick that all incident light is
absorbed or scattered before reaching the back surface of
Corresponding author at: CSIRO Industrial Physics, P.O. Box 218, the sample; typically a thickness of 1–3 mm is required. It
Lindfield, NSW 2070, Australia. Tel.: +61 294137150;
may also be reasonable to apply this method to coatings
fax: +16 294137200. that are sufficiently thick to absorb or scatter all the
E-mail address: tony.murphy@csiro.au incident light. Alternatively, for non-opaque coatings and

0927-0248/$ - see front matter Crown Copyright r 2007 Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.solmat.2007.05.005
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1327

substrates with optically smooth surfaces, a combination of the band gap of TiO2 sputtered onto indium–tin-oxide
reflectance and transmittance measurements allows the conducting glass from diffuse reflectance measurements.
optical properties of the coating to be determined by fitting Shankar et al. [23] used diffuse reflectance measurements of
of the data to standard optical equations that take into TiO2 nanotubes on a titanium substrate to determine the
account reflection and transmission at each interface band gap, finding an decrease in the band gap for nitrogen-
[6–10]. All these standard methods allow the determination doped TiO2. Chen et al. [33] derived ‘optical absorption
of the absorption coefficient as a function of wavelength, or thresholds’ of TiO2 and nitrogen-doped TiO2 films,
at least a variable proportional to the absorption coeffi- deposited on an aluminium substrate by reactive magne-
cient; it is then possible to use standard techniques to find tron sputtering, from diffuse reflectance measurements,
the band gap. The absorption coefficient also allows the finding an increase in the threshold wavelength for the
depth distribution of light absorption in the semiconductor doped films.
to be determined. One purpose of this paper is to investigate the validity of
Spectroscopic ellipsometry is also widely used to find the such attempts to derive band gaps of semiconductor
optical properties of thin film coatings with optically coatings from measurements of diffuse reflectance. For
smooth surfaces, but it generally does not allow accurate this purpose, I use a two-flux (Kubelka–Munk) radiative
determination of the absorption coefficient for wavelengths transfer model to investigate the absorption and scattering
at which the coating does not absorb strongly [11,12]. of the radiation in the semiconductor coating. The
In many cases of interest for water splitting, however, the Kubelka–Munk model is the basis for measurements of
above techniques cannot be applied. The semiconductor the band gap of thick powder samples. To apply the model
coating may be too thin to completely absorb or scatter the to thin coatings, reflections at the semiconductor–air and
incident radiation, and further the surfaces of the coating semiconductor–substrate interfaces are taken into account
and substrate may be optically rough, or the substrate may in a manner that is applicable to both optically smooth and
be opaque. For example, many workers have investigated optically rough surfaces. The dependence of the calculated
titanium dioxide coatings on a titanium substrate, formed diffuse reflectance on parameters such as the absorption
variously by thermal oxidation of a titanium sheet [13–19], coefficient, scattering coefficient, film thickness and surface
anodic electrodeposition on titanium sheet [13], ‘micro- roughness is investigated. The implications for efforts to
plasma’ oxidation of titanium sheet [20], magnetron determine the band gap are examined.
sputtering on a titanium mesh substrate [21], spin coating The second purpose of the paper is to demonstrate a
of TiO2 nanopowder onto titanium sheet [22], and growth method by which the band gap can be derived from
of arrays of TiO2 nanotubes on a titanium substrate measurements of diffuse reflectance. To this end, measure-
[23–25]. Examples of TiO2 coatings on different opaque ments of the diffuse reflectance of titanium dioxide films of
substrates include chemical vapour deposition of TiO2 onto different thicknesses on a titanium substrate are reported.
semiconductors [26], dip coating of TiO2 dip onto a gold The measurements are first used to check the calculated
foil substrate [27], and electrophoretic and spray coating of dependence of the diffuse reflectance on film thickness. An
TiO2 powder onto stainless steel as well as titanium [28]. inversion method, described elsewhere [34], is used to
Other semiconductor coatings include WO3 on a tungsten determine the absorption coefficient, and the band gap is
substrate, formed by both oxidation of tungsten sheet and then derived using standard methods. The inversion model
sputtering of WO3 powder onto the tungsten sheet [29], and is also applied to the measurements of Khan et al. [14] to
coatings of CdTe and CdS on titanium formed by cathodic estimate the absorption coefficient and band gap of their
electrodeposition, and CdS on titanium formed by carbon-doped TiO2 photoelectrode. From this, the spectral
chemical bath deposition [30]. dependence of the depth of absorption of light in the
In some of these cases, authors have attempted to derive semiconductor is calculated, and the visible light contribution
band-gap energies and draw conclusions about absorption to the water-splitting efficiency of the photoelectrode is
from measurements of diffuse reflectance, without calculat- estimated. The implications for visible water-splitting efficiency
ing the absorption coefficient. Khan et al. [14] used of anion-doped TiO2 photoelectrodes in general are discussed.
measurements of the diffuse reflectance of carbon-doped The paper is arranged as follows. Theoretical and
rutile TiO2, formed by oxidizing titanium sheet in a experimental methods are described in Section 2. In
methane flame, to deduce a band-gap wavelength of Section 2.1, the Kubelka–Munk radiative transfer model
535 nm, much larger than that derived for undoped rutile, and its application to diffuse reflectance measurements are
and to hypothesize that the TiO2 absorbed strongly up to outlined. Reflection coefficients from optically rough sur-
this wavelength. So et al. [31] used diffuse reflectance faces, required in the Kubelka–Munk expressions for diffuse
measurements of CdS–TiO2 particulate films to examine reflectance, are presented in Section 2.2. The method used to
the influence of TiO2 particles on the band gap. Shangguan produce the TiO2 coatings, and the properties of these
et al. [32] measured the diffuse reflectance of 0.3% Pt/TiO2 coatings, are given in Section 2.3. The measurements of
films on glass, and commented that the thinnest film had a diffuse reflectance are described in Section 2.4.
larger band gap than the other films, even though its Theoretical results obtained by application of the
photocatalytic activity was the same. He et al. [21] derived Kubelka–Munk model to TiO2 coatings on a titanium
ARTICLE IN PRESS
1328 A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337

substrate are presented in Section 3. The dependence of where the average crossing parameter e is defined such that
diffuse reflectance on coating properties is examined in the average path length travelled by diffuse light crossing a
Section 3.1, and the influence of the coating thickness on length dz is e dz. For collimated light, e ¼ 1, while for
the apparent band gap of the coating, derived directly from semiisotropic (i.e., isotropic in the direction of propaga-
the wavelength dependence of the diffuse reflectance, is tion) diffuse light, e ¼ 2 [39]. It is usual in applying the
considered in Section 3.2. Kubelka–Munk model to write z ¼ 12 and e ¼ 2, so that
Diffuse reflectance measurements of the TiO2 coatings S ¼ s and K ¼ 2k.
are presented and discussed in Section 4. The measure- In applying the Kubelka–Munk model to real coatings,
ments are used to examine two methods of determination it is important to take into account reflection from the
of the band gap; directly from the diffuse reflectance in surfaces and interfaces. Here a coating on an opaque
Section 4.1, and via inversion of the diffuse reflectance to substrate under collimated illumination is considered. The
obtain the absorption coefficient in Section 4.2. light reflected from the front surface may have both
In Section 5, the implications of these results for water collimated and diffuse components, but it is assumed that
splitting using doped electrodes that absorb visible light are the light transmitted into the coating is diffuse. As noted
considered, using the particular example of the diffuse above, this is in most cases reasonable in the case of a
reflectance measurements of carbon-doped TiO2 electrodes rough surface at the air–coating interface.
presented by Khan et al. [14]. Conclusions are presented in Note that the term scattering is sometimes used to
Section 6. describe what is here called diffuse reflection. It is also
important to distinguish between reflectance, which refers
to the reflection of light from the complete coating–sub-
2. Methods strate system, and reflection coefficients, which refer to
reflection from a single surface or interface. I will use R to
2.1. Kubelka–Munk model denote reflectance, and r to denote a reflection coefficient.
The following reflection coefficients have to be considered:
The Kubelka–Munk model [35,36] allows calculation of rfcc and rfcd , which denote coefficients for the reflection of
reflectance from a layer that both scatters and absorbs collimated light from the front of the air–coating interface,
light. It is a ‘two-flux’ model, which means that only diffuse as collimated light and as diffuse light, respectively; and rbdd
light is considered. This is appropriate where strong and rsdd , which denote coefficients for the reflection of
scattering occurs, where the incident light is diffuse, or diffuse light as diffuse light, from the back of the
where surfaces are optically rough [37]. In cases where the air–coating interface, and from the front of the coat-
light is not fully diffuse, four-flux models [38], which treat ing–substrate interface, respectively.
both collimated and diffuse light, are more appropriate. In The diffuse reflectance in terms of the reflection
standard spectrophotometers, such as that used here, the coefficients from the different interfaces, the coating
incident light is collimated, and in many coatings, thickness, and the scattering and absorption coefficients
scattering is weak. However, for coatings produced by is then given by [37]
oxidation of metals, by sol-gel techniques, and by some   
other means, the surface of the coating is optically rough, f 1  rfcd  rfcc 1  rbdd RKM
Rcd ¼ rcd þ , (3)
and it is reasonable to use the Kubelka–Munk model. 1  rbdd RKM
The Kubelka–Munk model uses an effective scattering
coefficient S and an effective absorption coefficient K to where
describe the optical properties of the coating. The effective 1  rsdd ½a  b cothðbShÞ
scattering coefficient is related to the usual scattering RKM ¼ . (4)
a þ b cothðbShÞ  rsdd
coefficient s by
S ¼ 2ð1  zÞs, (1) a ¼ ðS þ K Þ=S, (5)
pffiffiffiffiffiffiffiffiffiffiffiffiffi
where the forward scattering ratio z is defined as the ratio b¼ a2  1 , (6)
of the energy scattered by a particle in the forward
hemisphere to the total scattered energy. For Rayleigh and h is the thickness of the coating. The collimated
scattering, z ¼ 12, while for Mie scattering, 12ozo1. The reflectance is given simply by
effective absorption coefficient is related to the usual Rcc ¼ rfcc . (7)
absorption coefficient k by
The depth distribution of the absorbed power, normalized
K ¼ k, (2) to the input light flux, is given by [34]

      
1  rfcc  rfcd b 1 þ brsdd coshðSbh  SbzÞ þ ð1 þ aÞ 1  rsdd sinhðSbh  SbzÞ
að z Þ ¼ K     , (8)
b 1  rbdd rsdd coshðSbhÞ þ a  rbdd  rsdd þ arbdd rsdd sinhðSbhÞ
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1329

where z is the distance from the surface of the coating; a(z) rate data given by Dechamps and Lehr [44], and the values
has dimension m1. The absorbed photon flux per unit of s0 and t of the surface, measured with an atomic force
distance (photons m3 s1) is then microscope, are given in Table 1. While s0 and t were
Z 1 measured for only two of the samples, it is reasonable to
J z ð zÞ ¼ I l al ðzÞ dl, (9) assume that they will have similar values for the other
0 sample. The TiO2 coatings produced appeared to be
where Il (photons m2 s1 nm1) is the incident spectral opaque.
photon flux in the wavelength band from l to l+dl, and
where a(z) is written al(z) to denote its value in the 2.4. Diffuse reflectance measurements
wavelength band from l to l+dl. The cumulative absorbed
photon flux (photons m2 s1) to depth d is given by The diffuse reflectance of the samples was measured
Z dZ 1 using a Cary 5 spectrophotometer fitted with a diffuse
J C ðd Þ ¼ I l al ðzÞ dl dz. (10) reflectance attachment. The incident beam is collimated,
0 0 and reflected light is captured by an integrating sphere. A
matt Teflon reference was used to provide a nominal 100%
2.2. Calculation of reflection coefficients for an optically- reflectance measurement.
rough surface
3. Kubelka–Munk results
Expressions for the reflection coefficients from the front
surface rfcc and rfcd were derived for perpendicular incidence 3.1. Dependence of reflectance on coating properties
of collimated light, assuming that the height distribution of
the surface is Gaussian and spatially isotropic, and Fig. 1 shows the diffuse reflectance Rcd , and the
therefore can be described by the rms surface roughness reflection coefficients rfcc , rfcd , rbdd and rsdd , calculated for
s0 and the autocorrelation length t [34,37]. The diffuse– the parameters h, s0 and t of sample 3. Also shown are the
diffuse reflection coefficients rbdd and rsdd were calculated literature values of the refractive index n and the extinction
using an average over all angles of incidence of the Fresnel coefficient k for rutile TiO2 that were used in the
reflection coefficient [34,37]. calculations. It is assumed in this instance that the
All the reflection coefficients depend on the complex scattering coefficient S is negligible. The expression used
refractive indices n+ik of the materials on either side of an for Rcd is given in Eq. (3); full details of the calculations
interface. Values of the refractive index n and the extinction were given in Ref. [37].
coefficient k for TiO2 were taken from Cardona and The reflection coefficients follow the general shape of the
Harbeke [40] and Devore [41] as reported by Ribarsky [42]. refractive index, with a peak at 320 nm. The diffuse
Values of n and k for titanium were taken from Lynch and reflectance also follows this general shape, but shows an
Hunter [43]. Note that the absorption coefficient is related additional feature, a dip at wavelengths around 390 nm,
to k by corresponding to the wavelength at which the extinction
coefficient, and therefore the absorption coefficient, ap-
k ¼ 4pk=l, (11)
proaches zero. The dip is a consequence of the transition
where l is the wavelength of the light. from Rcd  rfcd at wavelengths for which absorption is
strong, to Rcd 4rfcd at wavelengths for which the absorption
2.3. Preparation of TiO2 coatings on titanium substrates is weak. It is this dip in the diffuse reflectance that provides
information about the band gap of the TiO2 coating.
Rutile TiO2 coatings were produced by oxidizing a piece It is instructive to examine the influence of the various
of titanium sheet (Sigma Aldrich, 99.7%, 0.25 mm thick) in parameters on the diffuse reflectance. Fig. 2 shows the
oxygen at 1 bar in a tube furnace. The titanium sheet was effect of varying the refractive index n, the extinction
etched in Kroll’s solution (one part 40% HF, one part 70% coefficient k and the average crossing parameter e of
HNO3 and three parts water) for 10 s prior to oxidation, to the coating. The reflectance is strongly dependent on the
give a rough substrate surface. The oxidation time and refractive index at all wavelengths; this is a result of
temperature, the thickness h estimated using the oxidation the dependence of the reflection coefficients on n. The
dependence on the extinction coefficient occurs in two
Table 1 wavelength regions. At wavelengths shorter than about
Oxidation conditions and properties of rutile TiO2 coatings on Ti 350 nm, where kX0:05, changes in k affect the reflection
coefficients. At wavelengths around 400 nm, where k is
Sample Oxidation time Temperature h s0 t
(Min) (1C) (mm) (mm) (mm) approaching zero, changes in k affect the absorption of
light in the coating, and therefore the reflectance. The
1 4.8 750 0.5 0.583 6.49 average crossing parameter e also affects the reflectance in
2 19 750 1.0
this region; a decrease in e is equivalent to a decrease in k
3 10 850 2.0 0.571 6.48
according to Eq. (2).
ARTICLE IN PRESS
1330 A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337

Fig. 1. Reflection coefficients and diffuse reflectance, calculated using the complex refractive index of TiO2 (shown) on titanium, for a coating of thickness
1 mm and surface roughness parameters of sample 3.

Fig. 3. (a) Effect of varying scattering coefficient S on the diffuse


reflectance Rcd; results are given for S corresponding to close-packed
spheres of different radii. (b) Effect of varying TiO2 layer thickness h on
the diffuse reflectance Rcd. (c) Effect of varying surface roughness on the
Fig. 2. Effect of varying (a) refractive index n, (b) extinction coefficient k diffuse reflectance Rcd, collimated reflectance Rcc, and total reflectance
and (c) average crossing parameter e on the diffuse reflectance Rcd for a 1 Rcd+Rcc. Standard conditions in all cases are a 1 um thick TiO2 coating
um thick TiO2 coating on titanium with the surface roughness of sample 3. on titanium with the surface roughness of sample 3 and negligible
scattering coefficient.

Fig. 3 shows the effect of varying the scattering using the Mie scattering computer code BHMIE [45]. The
coefficient, the TiO2 coating thickness and the rms surface calculation assumes spherical particles with the complex
roughness on the reflectance. The scattering coefficient S is refractive index of rutile TiO2, closely packed in an air
calculated using Eq. (1), with values of s and z calculated environment. The maximum value of S under this
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1331

assumption is of order 107 m1. For comparison, cient). In the next subsection, using the example of the
Kb107 m1 for lp340 nm and Kb107 m1 for coating thickness, I will examine the consequences for
lX380 nm. In real coatings, the scattering coefficient, attempts to determine the band gap from measurements of
and hence the effect of scattering, will be smaller, since the diffuse reflectance.
particles will not be separate; in some coatings, including
the ones studied in the experiments reported here, 3.2. Calculation of the band gap from the diffuse reflectance
scattering is negligible. Nevertheless, it is clear that
scattering can have a significant influence on the reflectance Many workers [14,21,23,32,33] have attempted to obtain
for lX350 nm. the band gap of a semiconductor coating on an opaque
Fig. 3b shows the effect of altering the thickness h of the substrate directly from diffuse reflectance measurements.
TiO2 coating. Increasing h shifts the dip in the reflectance They fitted a line to the long-wavelength edge of the dip in
that occurs around 400 nm to longer wavelengths; as noted the diffuse reflectance (see Fig. 1); the band-gap wavelength
earlier, this dip is associated with the decrease in the was taken to be the intersection of this line with a
absorption coefficient to zero. This effect will be examined horizontal line corresponding to the maximum reflectance.
in detail in Section 3.2. In practice, this was done by calculating the absorbance
The surface roughness of the TiO2 coating can also
influence the reflectance, as shown in Fig. 3c. Decreasing A ¼ Rmax  Rcd , (12)
the rms roughness s0 to 10 nm for t/s010 leads to a where Rmax is the maximum value of Rcd for wavelengths
decrease in the diffuse reflectance, and an increase in the longer than that of the dip in the diffuse reflectance. This
collimated reflectance Rcc; the total reflectance remains sets the minimum absorbance to zero; the band-gap
approximately the same. For smaller values of t/s0, the wavelength is obtained by extrapolating the long-wave-
total reflectance can decrease significantly [37]. length edge of the peak in absorbance to this zero line.
Overall, it is clear from Figs. 2 and Fig. 3 that the diffuse Fig. 4a, taken from Khan et al. [14], illustrates this
reflectance depends strongly on a large number of method. Khan et al. obtained a band-gap wavelength of
parameters. In the region around 400 nm where the 414 nm (3.0 eV) for the undoped TiO2, and 535 nm
extinction coefficient decreases to zero, the refractive index, (2.32 eV) and 440 nm (2.82 eV) for the carbon-doped TiO2.
the average crossing parameter, the scattering coefficient, Fig. 4b shows the application of the same method to
the coating thickness and the surface roughness can all calculated absorbance curves, obtained using Eqs. (3) and
have an effect on the reflectance as significant as that of the (12), for TiO2 coatings of different thicknesses. It is clear
extinction coefficient (and hence the absorption coeffi- that the band-gap wavelength obtained using the method

Fig. 4. Determination of the band-gap wavelength from diffuse reflectance measurements. (a) Method used to fit measurements of absorbance by Khan et
al. [14]. (b) Fitting by the same method of the calculated absorbance of TiO2 coatings of different thicknesses.
ARTICLE IN PRESS
1332 A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337

depends strongly on the thickness of the coating, increasing Fig. 6 shows the band gaps obtained for the three
from 390 nm (3.18 eV) to 419 nm (2.95 eV) as the thickness samples. These are compared to the band gap obtained
increases from 100 nm to 10 mm. from absorbances calculated using the Kubelka–Munk
It is clear from Fig. 2 and Fig. 3 and the discussion in model, such as those shown in Fig. 4b, for a range of
Section 3.1 that the band gap calculated using this method coating thicknesses. The band gaps of the samples are a
will also depend on other parameters, such as scattering little lower than the calculations predict, and the depen-
coefficient, refractive index and average crossing para- dence on the coating thickness is not as strong. This is
meter. mainly a consequence of the variations in the refractive
It is interesting to note that the wavelength dependence index of the TiO2 coatings with thickness.
of the diffuse reflectance of gallium arsenide has been It has been shown in an earlier publication that the
investigated by Johnson and Tiedje [46] for the application refractive index of the thickest sample, sample 3, is closest
of optical band-gap thermometry. They found that the to the value used in the calculations [34]; this is the sample
apparent band gap derived from the diffuse reflectance for which best agreement with the calculations is obtained.
depended on thickness of the gallium arsenide and its The refractive index for the other samples is significantly
surface texture, necessitating compensation if the tempera- lower than that used in the calculations [34]. This may be
ture was to be accurately measured. due to the presence of a suboxide layer that is expected to
be formed close to the substrate for all coatings [44,47], and
4. Measurements and discussion which will be of more relative significance for the thinner
coatings. It can be seen from Fig. 2a that decreasing the
4.1. Diffuse reflectance and band-gap determination of TiO2 refractive index leads to a decrease in the apparent band
samples gap.
The measurements confirm that the apparent band gap
In order to examine the validity of the calculations, and obtained by extrapolating the peak in the absorbance to
to investigate the possibility of determining the band gap the zero line depends on the thickness of the sample. The
and absorption coefficient from diffuse reflectance mea- question remains whether it is possible to determine the
surements, three samples with TiO2 coatings of different real band gap from measurements of the diffuse reflec-
thicknesses were prepared as described in Section 2.3. The tance. This is addressed in the next subsection.
diffuse reflectance of the samples is shown in Fig. 5a. The The results obtained by Shangguan et al. [32], who
band gap is obtained using the method described in Section examined the diffuse reflectance of 0.3% Pt/TiO2 films on
3.2; this is shown in Fig. 5b. glass, and found that the thinnest film had a lower

Fig. 5. (a) Measured diffuse reflectance for three samples with TiO2 coatings of different thicknesses. (b) Absorbance of TiO2 coatings, showing linear
extrapolations to obtain the band-gap wavelength.
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1333

Fig. 6. Dependence of apparent band gap energy, obtained by extrapolation of the peak in the calculated absorbance, on TiO2 coating thickness, and
band gaps for the three samples obtained by the same method from the measured absorbance.

Fig. 7. a) Absorption coefficient of the three samples derived from the diffuse reflectance measurements, and (b) Tauc plots showing possible fits to obtain
the band gap.

band-gap wavelength (i.e., higher band-gap energy) than the Kubelka–Munk expression for the diffuse reflectance
the other thicker films, even though its photocatalytic (Eq. (3)), thereby obtaining values of n and k (or
activity was the same, is consistent with the trend shown in equivalently, k) that give the best fit to the measurements.
Fig. 6. It is hence likely that the decreased band-gap The surface roughness parameters s0 and t, and the
wavelength for the thinnest film is apparent rather than thickness h were fixed, and the results were consistent with
real. a negligibly small scattering coefficient. Fig. 7a shows the
best fit values of the absorption coefficient for the three
4.2. Determination of band gap from absorption coefficient samples.
For crystalline solids with an indirect band gap, such as
It has been shown in an earlier publication [34] that it is rutile TiO2, the dependence of the absorption coefficient k
possible to derive the absorption coefficient from measure- on the frequency n can be approximated as
ments of the diffuse reflectance of the samples. This was  2
done using the spectral projected gradient method to invert khn ¼ A hn  E g , (13)
ARTICLE IN PRESS
1334 A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337

where A is a constant [48]. It is clear from Eq. (13) that the shows the resulting diffuse reflectance curves. The undoped
band gap Eg can be obtained by extrapolating to zero a TiO2 curve matches that for sample 3 closely, giving some
linear fit to a plot of (khv)1/2against hv (often referred to as confidence in the procedure.
a Tauc plot). Note that some authors exclude the hv factor The absorption coefficient was then obtained for the
on the left-hand side of Eq. (13) [49,50]; this has only a carbon-doped TiO2 case using the inversion method. The
minor influence on the value of Eg obtained. Fig. 7b shows thickness of the coating was set to 2 mm, as for sample 3.
Tauc plots for the three samples. A straight line fits the The result is shown in Fig. 8b. For wavelengths below
sample 3 data well, yielding a band gap of 3.0070.02 eV. about 400 nm, the absorption coefficient of the carbon-
The fits for samples 2 and 3 are not as good, so the doped TiO2 is similar to that of sample 3, reaching about
uncertainties in the band gap are larger; the band gaps are 108 m1. For longer wavelengths, the carbon-doped TiO2
estimated to be in the range 3.070.1 eV. shows only weak absorption, with absorption coefficient
These results indicate that it is possible to determine the between 104 and 105 m1 up to a wavelength of about
absorption coefficient, and hence the band gap, from 500 nm. Using Eq. (13), it is possible to derive two band
inversion of the measurements of diffuse reflectance, gaps, as shown in Fig. 9, corresponding to the main (sub-
although precise determination of the band gap may not 400 nm) section of the absorption coefficient against
always be possible. wavelength curve and the high-wavelength (400–510 nm)
section. The band-gap energies are 3.1370.02 and
5. Implications for water splitting by doped TiO2 electrodes 2.4470.02 eV, respectively. Both band gaps are larger than
the values of 2.82 and 2.32 eV, respectively, obtained by
In this section, I apply the methods that were used to Khan et al. [14] by fitting a line directly to the absorbance.
determine the absorption coefficient of the oxidized In solar water-splitting applications, only charge carriers
titanium samples to the diffuse reflectance measurements produced by absorption of photons within a relatively
of Khan et al. [14]. Unfortunately the measurements, small distance of the surface will reach the electrolyte
reproduced in Fig. 4a, are in arbitrary units and no zero is before recombining. This distance is material dependent,
given. The following procedure was adopted to circumvent and depends on the depth of the depletion layer and the
this problem. First, the diffuse reflectance was obtained diffusion length of the charge carriers. For a flame-oxidized
from the absorbance using Eq. (12) with Rmax ¼ 0. The TiO2 coating, it was measured to be about 200 nm [3]. It is
undoped TiO2 reflectance curve was then scaled and shifted important, therefore, to determine the depth profile of
vertically so that it matched the diffuse reflectance of photon absorption for the carbon-doped TiO2, which can
sample 3 at 389 nm (the wavelength at which the reflectance be calculated using Eq. (10), assuming illumination by the
of sample 3 is minimum) and 430 nm (the maximum standard AM1.5 global solar spectrum [51]. The results are
wavelength for which Khan et al. presented data). Fig. 8a shown in Fig. 10. Results are given for the full solar

Fig. 8. a) Diffuse reflectance of undoped and carbon-doped TiO2 coatings, modified from the data given by Khan et al. [14] so that the undoped TiO2
reflectance agrees with that of sample 3. (b) Absorption coefficients of the carbon-doped TiO2 coating and sample 3, obtained by inversion of the
reflectance in (a).
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1335

Fig. 9. Tauc plot of data of Khan et al. [14], showing fits to obtain the band gaps.

Fig. 10. Cumulative absorbed photon flux JC(d) versus depth d for the carbon-doped TiO2 coating of Khan et al. [14]. Results are given for the full AM1.5
global solar spectrum, and the sub-410 nm part of the spectrum. The ratio of the two results is also given.

spectrum, and the part of the spectrum below 410 nm they provide only an upper limit to the contribution of the
(corresponding to the 3.0 eV band gap of rutile TiO2). visible light, since absorbed photons do not necessarily
About 35% of the photons absorbed in the full 2mm depth result in water-splitting reaction, due to for example
of the coating are of wavelength longer than 410 nm, recombination of electron–hole pairs at trap states [52].
indicating that additional absorption due to the carbon- Khan et al. [14] claimed that their carbon-doped
doping of the TiO2 is significant. However, only 10% of the electrode had a water-splitting efficiency of 8.35% under
photons absorbed in the top 200 nm of the coating are of illumination by a xenon lamp with spectrum similar to
wavelength longer than 410 nm, indicating that the carbon AM1.5, compared with 1.08% for the undoped electrode.
doping of the TiO2 will only make a small contribution to Murphy et al. [3] and Hägglund et al. [53] have shown
water splitting. Allowing a larger distance of 500 nm that the 8.35% efficiency is not possible for the band
increases the percentage of absorbed photons that have gap of 2.32 eV (535 nm), even with 100% utilization of all
wavelength longer than 410 nm, but only to 17%. It is incident photons with energies greater than 2.32 eV in
emphasized that the calculations on which these figures are water-splitting reactions. Here it is shown that only a
based take into account the increased optical path length small increase in the efficiency is expected due to carbon-
due to the surface roughness of the electrode and multiple doping. This is consistent with the results obtained by
reflections; in the Kubelka–Munk model, it is assumed that other workers [18,19,52] who have been unable to
the light is diffuse, and reflections from the air–coating and reproduce Khan et al.’s efficiencies for carbon-doped
coating–substrate interfaces are taken into account. Also, TiO2 electrodes.
ARTICLE IN PRESS
1336 A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337

It should be noted that because Khan et al. [14] provided destruction of organic molecules [54–56]. However, these
neither a scale for their absorbance data nor thicknesses of studies have used suspensions of doped TiO2 powder, in
their coatings, the above calculations cannot be precise. which case very large optical depths are obtainable. In such
Nevertheless, the conclusions that the absorption coeffi- applications, the weak visible absorption coefficient is far
cient in the visible light region is orders of magnitude less important than in photoelectrochemical water-split-
smaller than in the ultraviolet region, and that the ting, in which a solid photoelectrode is required.
absorption of visible light will make only a minor
contribution to the water-splitting efficiency, are robust. 6. Conclusions
This is a consequence of the fact that a small decrease in
absorbance corresponds to a decrease of orders of The Kubelka–Munk radiative transfer model, modified
magnitude in the absorption coefficient, and the require- to treat optically rough surfaces, has been applied to TiO2
ment that photons be absorbed within about 200 nm of the coatings on a titanium substrate, as used for solar water
surface to be effective in water splitting. splitting. It has been shown that the diffuse reflectance
These considerations can also be applied to the diffuse depends on parameters including the coating thickness,
reflectance measurements of other workers. For example, refractive index, scattering coefficient and surface rough-
Shankar et al. [23] found an increased absorbance at visible ness, as well as the absorption coefficient. Hence, deriva-
wavelengths for nitrogen-doped TiO2 nanotubes compared tion of the absorption coefficient and band gap of the
to undoped TiO2 nanotubes. However, the absorbance at coating has to take into account all these parameters. The
visible wavelengths is 50% or less of that at ultraviolet effect of the coating thickness has been examined in detail,
wavelengths, so the absorption coefficient will be greatly and it has been shown that the widely used practice of
decreased. It should be noted here that quantitative determining the band gap of the coating directly from the
analysis is more difficult in this case because of the diffuse reflectance is likely to be inaccurate. However, it has
complicated geometry. been found that, by applying an inversion technique to the
It is possible to mitigate the effects of a low absorption diffuse reflectance measurements, it is possible to derive the
coefficient by employing nanostructured electrodes, incor- absorption coefficient, and hence the band gap, more
porating for example large-aspect ratio nanotubes or reliably.
nanorods aligned perpendicular to the substrate. Using This inversion technique has been also applied to the
such structures, electrode thicknesses parallel to the measurements of diffuse reflectance of carbon-doped TiO2
incident light can be increased significantly, while photons electrodes presented by Khan et al. [14], who measured an
are still absorbed close to the surface. Nevertheless, as extension of the absorption into the visible light region due
noted above, Fig. 10 shows that even with a 2 mm electrode to carbon doping. It is found that the band gap estimated
thickness, only 35% of the absorbed photons are at by Khan et al. is inaccurate, and further that the
wavelengths longer than 410 nm.The theoretical maximum absorption coefficient in the visible region is orders of
efficiency for a semiconductor absorbing all photons at magnitude lower than in the ultraviolet region, with the
wavelengths shorter than 410 nm is 2.2% [3], so the consequence that only a small proportion of the visible
maximum efficiency for a 2 mm thick carbon-doped TiO2 photons are absorbed near the surface of the electrode. As
nanostructured electrode with the absorption coefficient a consequence, the visible photons are expected to make
shown in Fig. 9b is 3.0%. This neglects losses due to only a small contribution to the water-splitting efficiency.
reflection, the requirement for a bias voltage, etc. Use of electrodes incorporating large-aspect-ratio nanos-
Recent experimental studies have investigated water- tructures, such as nanotubes or nanorods, may be a means
splitting efficiencies for carbon-doped TiO2 nanotube of mitigating the effects of the low visible absorption
arrays [24,25]. One of the studies showed an increase in coefficients.
efficiency with nanotube length for undoped nanotubes
[25], and both demonstrated significant visible activity for Acknowledgements
doped nanotubes, although overall efficiencies were still
below 1%, in accordance with the calculation above. In I thank Dr Piers Barnes for providing rms roughness and
one case, incident photon conversion efficiency (IPCE) autocorrelation length data for the TiO2 samples, and Dr
measurements were used to measure the band gap using a Barnes, Ms Julie Glasscock and Dr Ian Plumb for useful
Tauc plot [25]. This method, which is widely used, requires discussions and comments.
the assumption that the absorption coefficient for inter-
band transitions is proportional to the photocurrent References
density, but does have the advantage that absorption due
to transitions not related to the band gap, such as the d–d [1] M.F. Weber, M.J. Dignam, J. Electrochem. Soc. 131 (1984)
transitions that occur in transition-metal-doped oxides, are 1258–1265.
[2] J.R. Bolton, S.J. Strickler, J.S. Connolly, Nature 316 (1985) 495–500.
ignored. [3] A.B. Murphy, P.R.F. Barnes, L.K. Randeniya, I.C. Plumb, I.E.
A number of other recent studies have found visible Grey, M.D. Horne, J.A. Glasscock, Int. J. Hydrogen Energy 31
photocatalytic activity using carbon-doped TiO2 in the (2006) 1999–2017.
ARTICLE IN PRESS
A.B. Murphy / Solar Energy Materials & Solar Cells 91 (2007) 1326–1337 1337

[4] B. Karvaly, I. Hevesi, Z. Naturforsch. 26a (1971) 245–249. [31] W.-W. So, K.-J. Kim, S.-J. Moon, Int. J. Hydrogen Energy 29 (2004)
[5] W.W. Wendlandt, H.G. Hecht, Reflectance Spectroscopy, Wiley 229–234.
Interscience, New York, 1966. [32] W. Shangguan, A. Yoshida, M. Chen, Solar Energy Mater. Solar
[6] O.S. Heavens, Optical Properties of Thin Solid Films, Butterworths, Cells 80 (2003) 433–441.
London, 1955. [33] S.-Z. Chen, P.-Y. Zhang, D.-M. Zhuang, W.-P. Zhu, Catal.
[7] E. Shanti, E. Dutta, A. Banerjee, K.L. Chopra, J. Appl. Phys. 51 Commun. 5 (2004) 677–680.
(1981) 6243–6251. [34] A.B. Murphy, Appl. Opt., doc. 46 (2007) 3133–3143.
[8] S.P. Lyashenko, V.K. Miloslavskii, Opt. Spectrosc. 16 (1964) 80–81. [35] P. Kubelka, F. Munk, Z. Tech. Phys. 12 (1931) 593–601.
[9] R. Swanepoel, J. Phys. E: Sci. Instrum. 16 (1983) 1214–1222. [36] P. Kubelka, J. Opt. Soc. Am. 38 (1948) 448–457.
[10] K. L. Eskins and K. W. Mitchell, in: Proceedings of the 18th IEEE [37] A.B. Murphy, J. Phys. D: Appl. Phys. 39 (2006) 3571–3581.
Photovoltaic Specialists Conference, (IEEE, 1985), pp. 720–725. [38] B. Maheu, J.N. Letoulouzan, G. Gouesbet, Appl. Opt. 23 (1984)
[11] D.E. Aspnes, A.A. Studna, Phys. Rev. B 27 (1983) 985–1009. 3353–3362.
[12] G.E. Jellison, L.A. Boatner Jr., J.D. Budai, B.-S. Jeong, D.P. Norton. [39] G. Kortüm, Reflectance Spectroscopy, Springer, New York, 1969.
J. Appl. Phys. 93 (2003) 9537–9541. [40] M. Cardona, G. Harbeke, Phys. Rev. 137 (1965) A1467–A1476.
[13] A. Fujishima, K. Kohayakawa, K. Honda, J. Electrochem. Soc. 122 [41] J.R. Devore, J. Opt. Soc. Am. 41 (1951) 416–419.
(1975) 1487–1489. [42] M.W. Ribarsky, Titanium dioxide (TiO2) (rutile), in: E.D. Palik
[14] S.U.M. Khan, M. Al-Shahry, W.B. Ingler Jr., Science 297 (2002) (Ed.), Handbook of Optical Constants, Academic, Orlando, Florida,
2243–2245. 1985, pp. 795–800.
[15] J. Akikusa, S.U.M. Khan, Int. J. Hydrogen Energy 22 (1997) [43] D.W. Lynch, W.R. Hunter, Introduction to the data for several
875–882. metals, in: E.D. Palik (Ed.), Handbook of Optical Constants of
[16] S.-E. Lindquist, A. Lindgren, Y.-N. Zhu, J. Electrochem. Soc. 132 Solids, vol. III, Academic, Orlando, Florida, 1998, pp. 233–286.
(1985) 623–631. [44] M. Dechamps, P. Lehr, J. Less-Common Metals 56 (1977)
[17] S.U.M. Khan, J. Akikusa, J. Electrochem. Soc. 145 (1998) 89–93. 193–207.
[18] P.R.F. Barnes, L.K. Randeniya, A.B. Murphy, P.B. Gwan, I.C. [45] C.F. Bohren, D.R. Huffman, Absorption and Scattering of Light by
Plumb, J.A. Glasscock, I.E. Grey, C. Li, Dev. Chem. Eng. Mineral Small Particles, Wiley-Interscience, New York, 1983.
Process. 14 (2006) 51–70. [46] S.R. Johnson, T. Tiedje, J. Crystal Growth 175/176 (1997) 273–280.
[19] K. Noworyta, J. Augustynski, Electrochem. Solid State Lett. 7 (2004) [47] P.R.F. Barnes, L.K. Randeniya, P.F. Vohralik, I.C. Plumb,
E31–E33. J. Electrochem. Soc. 154 (2007) H249–H257.
[20] X. Wu, Z. Jiang, H. Liu, S. Xin, X. Hu, Thin Solid Films 441 (2003) [48] E.J. Johnson, Absorption near the fundamental edge, in: R.K.
130–134. Willardson, A.C. Beers (Eds.), Optical Properties of III-V Com-
[21] C. He, X.Z. Li, N. Graham, Y. Wang, Appl. Catal. A: General 305 pounds, Semiconductors and Semimetals, vol. 3, Academic Press,
(2006) 54–63. New York, 1967, pp. 153–258.
[22] P.R. Mishra, P.K. Shukla, A.K. Singh, O.N. Srivastava, Int. J. [49] J.I. Pankove, Optical Processes in Semiconductors, Dover, New
Hydrogen Energy 28 (2003) 1089–1094. York, 1971.
[23] K. Shankar, K.C. Tep, G.K. Mor, C.A. Grimes, J. Phys. D: Appl. [50] C.R. Aita, Y.-L. Liu, M.L. Kao, S.D. Hansen, J. Appl. Phys. 60
Phys. 39 (2006) 2361–2366. (1986) 749–753.
[24] K. Shankar, M. Paulose, G.K. Mor, O.K. Varghese, C.A. Grimes, [51] ASTM, Standard tables for reference solar spectral irradiance at air
J. Phys. D.: Appl. Phys. (2005) 3543–3549. mass 1.5: Direct normal and hemispherical for a 371 tilted surface,
[25] J.H. Park, S. Kim, A.J. Bard, Nano Lett 6 (2006) 24–28. American Society for Testing and Materials Standard G159-98, West
[26] P.A. Kohl, S.N. Frank, A.J. Bard, J. Electrochem. Soc. 124 (1977) Conshohocken, Pennsylvania, 1998.
225–229. [52] B. Neumann, P. Bogdanoff, H. Tributsch, S. Sakthivel, H. Kisch,
[27] W. Sun, C.R. Chenthamarakshan, K. Rajeshwar, J. Phys. Chem. B J. Phys. Chem. B 109 (2005) 16579–16586.
106 (2002) 11531–11538. [53] C. Hägglund, M. Grätzel, B. Kasemo, Science 301 (2003)
[28] J.A. Byrne, B.R. Eggins, N.M.D. Brown, B. McKinney, M. Rouse, 1673b.
Appl. Catal. B: Environ. 17 (1998) 25–36. [54] S. Sakthivel, H. Kisch, Angew. Chem. Int. Ed. 42 (2003) 4908–4911.
[29] W. Gissler, R. Memming, J. Electrochem. Soc. 124 (1977) 1710–1714. [55] H. Irie, Y. Watanabe, K. Hashimoto, Chem. Lett. 32 (2003) 772–773.
[30] R.K. Pandey, S. Mishra, S. Tiwari, P. Sahu, B.P. Chandra, Solar [56] C. Xu, R. Killmeyer, M.L. Grey, S.U.M. Khan, Appl. Catal. B:
Energy Mater. Solar Cells 60 (2000) 59–72. Envir. 64 (2006) 312–317.

Anda mungkin juga menyukai