You are on page 1of 8

Applied Catalysis A: General 400 (2011) 195–202

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Visible light responsive iodine-doped TiO2 for photocatalytic reduction of


CO2 to fuels
Qianyi Zhang a , Ying Li a,∗ , Erik A. Ackerman b , Marija Gajdardziska-Josifovska c , Hailong Li d
a
University of Wisconsin-Milwaukee, Department of Mechanical Engineering, Milwaukee, WI 53211, USA
b
University of Wisconsin-Milwaukee, Department of Electrical Engineering, Milwaukee, WI 53211, USA
c
University of Wisconsin-Milwaukee, Department of Physics and Laboratory for Surface Studies, Milwaukee, WI 53211, USA
d
University of Florida, Department of Environmental Engineering and Sciences, Gainesville, FL 32611, USA

a r t i c l e i n f o a b s t r a c t

Article history: Iodine-doped titanium oxide (I-TiO2 ) nanoparticles that are photocatalitically responsive to visible light
Received 28 February 2011 illumination have been synthesized by hydrothermal method. The structure and properties of I-TiO2
Received in revised form 21 April 2011 nanocrystals prepared with different iodine doping levels and/or calcination temperatures were char-
Accepted 25 April 2011
acterized by X-ray diffraction, transmission electron microscopy and diffraction, X-ray photoelectron
Available online 3 May 2011
spectroscopy, and UV–vis diffuse reflectance spectra. The three nominal iodine dopant levels (5, 10,
15 wt.%) and the two lower calcination temperatures (375, 450 ◦ C) produced mixture of anatase and
Keywords:
brookite nanocrystals, with small fraction of rutile found at 550 ◦ C. The anatase phase of TiO2 increased
Photocatalysis
Solar energy conversion
in volume fraction with increased calcination temperature and iodine levels. The photocatalytic activi-
TiO2 ties of the I-TiO2 powders were investigated by photocatalytic reduction of CO2 with H2 O under visible
Nanocomposite light ( > 400 nm) and also under UV–vis illumination. CO was found to be the major photoreduction
CO2 reduction product using both undoped and doped TiO2 . A high CO2 reduction activity was observed for I-TiO2 cata-
lysts (highest CO yield equivalent to 2.4 ␮mol g−1 h−1 ) under visible light, and they also had much higher
CO2 photoreduction efficiency than undoped TiO2 under UV–vis irradiation. I-TiO2 calcined at 375 ◦ C has
superior activity to those calcined at higher temperatures. Optimal doping levels of iodine were identi-
fied under visible and UV–vis irradiations, respectively. This is the first study that investigates nonmetal
doped TiO2 without other co-catalysts for CO2 photoreduction to fuels under visible light.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction promising candidate because of its strong redox ability, low cost,
stability, and environmental benignness. However, one challenge
The global warming effect is believed to be associated with for the application of semiconductor photocatalysts like TiO2 is the
the increasing concentrations of greenhouse gases in the atmo- fast recombination of photo-induced holes (h+ ) and electrons (e− ).
sphere, where the major contribution comes from CO2 emissions Another challenge is the requirement of ultraviolet (UV) light exci-
from fossil fuel consumption. Current pre- and post-combustion tation due to the wide band gap of TiO2 (3.2 eV for anatase and
CO2 capture and sequestration technologies are energy intensive 3.0 eV for rutile). As a result, the efficiency of CO2 conversion to
and thus costly [1]. Furthermore, there are many uncertainties fuels is generally low.
with regard to long-term storage of CO2 in geological formations Modification of TiO2 with metal (e.g. Pt, Pd, Ag, and Cu) parti-
[1]. In contrast, with recent innovations in photocatalysis, recy- cles or clusters has been reported to inhibit charge recombination
cling CO2 to fuels using sunlight as the sole energy input offers possibly because the metals serve as electron traps [14]. Thus, an
a brand new opportunity for a sustainable energy future. This can increased CO2 photoreduction efficiency was observed for metal
not only mitigate CO2 emissions but also produce energy-bearing modified TiO2 . Tseng et al. [3] synthesized Cu/TiO2 catalysts by
compounds such as CO, methane, and methanol [2–4] that can a sol–gel method and found the rate of CO2 photoreduction to
be subsequently converted to liquid transportation fuels. Materi- methanol was much higher than those without copper loading.
als that have been reported for CO2 photoreduction applications Li et al. [13] reported markedly increased CO2 photo-conversion
include ZrO2 [5,6], MgO [7], NiO/InTaO4 [8], Ga2 O3 [9], photosen- efficiency by Cu/TiO2 catalyst dispersed on mesoporous silica
sitized complexes [10], and TiO2 based catalysts [11–13]. TiO2 is a and selective CH4 production due to Cu loading. However, too
high a concentration of metal dopant may form recombination
centers that lead to a reduced photocatalytic efficiency. Opti-
∗ Corresponding author. Tel.: +1 414 229 3716; fax: +1 414 229 6958. mal metal concentrations have been reported for modified TiO2
E-mail address: liying@uwm.edu (Y. Li). (e.g. with Ag or Cu) for both photooxidation and photoreduction

0926-860X/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2011.04.032
196 Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202

applications [13,15–17]. While metal modifications on TiO2 have ing 3 ml of titanium isopropoxide (TTIP) (Acros Organic, >98%) in
apparent enhancement in charge separation, they have limited con- 3 ml of anhydrous isopropanol (Acros Organic, >99.5%). The mix-
tribution to extending the photo-response to visible light region. ture was then added dropwise into a solution of iodic acid (HIO3 )
Sasirekha et al. [18] observed that Ru doped TiO2 has almost the (Alfa Aesar, >99.5%) with continuous stirring for 2 h. After the reac-
same absorption spectra as the undoped TiO2 . Dholam and Patel tion, the resultant white mixture was transferred to a Teflon-lined
[19] reported that Cr and Fe doped TiO2 prepared by a sol–gel vessel for hydrothermal treatment at 100 ◦ C for 12 h. The resultant
method had a very limited effect on inducing a red-shift in TiO2 yellow particles were filtrated and washed with copious amount of
absorption spectra compared to those prepared by a magnetron de-ionized water until pH 7 followed by drying in an oven at 80 ◦ C
sputtering method. for 1 h. The samples were finally calcined in air for 2 h at different
On the other hand, it has been widely reported that doping or temperatures (375, 450, or 550 ◦ C). Different iodine doping levels
co-doping TiO2 with nonmetals (e.g. C, N, S, F, etc.) has resulted in were prepared by varying the quantity of HIO3 added (0.06–0.18 g).
more significant band gap narrowing compared to metal doping, The samples are denoted in the way of “x% I-TiO2 -yC”, where x is
leading to high photocatalytic efficiency under visible light irra- the nominal weight percentage of iodine in the sample (calculated
diation [20–27]. Wu et al. [26] reported that the band gaps of N from the bulk solution) and y is the calcination temperature. For
doped and N–B co-doped TiO2 were 2.16 eV and 2.13 eV, respec- example, 5% I-TiO2 -375C represents 5 wt.% (nominal) I-doped TiO2
tively, much smaller than that of pure TiO2 (3.18 eV for anatase). calcined at 375 ◦ C. For comparison, undoped TiO2 was prepared
Pelaez et al. [27] synthesized N–F co-doped TiO2 that exhibited following the same procedure without adding HIO3 . All samples
high surface area, low degree of agglomeration and high activity were grinded and sieved by a 45 ␮m stainless steel sieve before
in degradation of microcystin under visible light. Recently, less characterization and photoreduction experiments.
studied iodine has been doped into TiO2 , and improved visible
light activity towards the decomposition of organic compounds
2.2. Photocatalyst characterization
has been reported [28–30]. The structural and electronic proper-
ties of I-doped TiO2 were investigated based on density functional
Brunauer–Emmett–Teller (BET) surface area analysis by N2
theory (DFT) calculations, and the results indicated that substi-
adsorption was performed using a Quantachrome NOVA 1200 gas
tutional iodine contributes to a much more efficient and stable
sorption analyzer (Boynton Beach, FL). The crystal structures of the
photocatalyst than pristine TiO2 [31]. In comparison to other non-
prepared catalysts were identified by X-ray diffraction (XRD) (Scin-
metal dopants (N, C, B, S), iodine doping may result in superior
tag XDS 2000) using Cu K␣ irradiation at 45 kV and a diffracted
photocatalytic activity due to the following reasons. First, unlike
beam monochromator operated at 40 mA in the 2 range from 20◦
other nonmetal dopants that substitute lattice oxygen, iodine was
to 70◦ at a scan rate of 1◦ /min. The crystal size of different crys-
reported to be able to replace lattice titanium due to the close
tal phases was calculated by the Scherrer equation. The fractional
ionic radii of I5+ and Ti4+ [28,32]. The substitution of Ti4+ with I5+
phase content, WA , WB , and WR , for anatase, brookite, and rutile,
causes charge imbalance and results in the generation of Ti3+ sur-
respectively, are mathematically defined in Eqs. (1)–(3) [34]:
face states that may trap the photoinduced electrons and forestall
charge recombination [32]. In addition, first principle calculations 0.886 × AA
suggest that iodine atoms prefer to be doped near the TiO2 surface WA = (1)
(0.886 × AA + AR + 2.721 × AB )
due to the strong I–O repulsion [29], and thus, the surface doped
I5+ will not only trap electrons but also facilitate electron transfer 2.721 × AB
WB = (2)
to the surface adsorbed species [29,33]. Finally, it is suggested that (0.886 × AA + AR + 2.721 × AB )
the continuous states consisting of 5p and/or 5s orbitals of I5+ and
AR
O 2p orbitals of the valence band are favorable for efficient trap- WR = (3)
ping of holes at the I-induced states in the TiO2 particle (not on the (0.886 × AA + AR + 2.721 × AB )
surface), which causes a decrease in the oxidation power [33]. For where AA , AB and AR represent the integrated intensity of
CO2 photoreduction, one of the challenges is the re-oxidation of the the anatase (1 0 1) peak (2 = 25.28◦ ), the brookite (1 2 1) peak
CO2 reduction products by h+ or OH• radicals. Hence, the impaired (2 = 30.81◦ ), and the rutile (1 1 0) peak (2 = 27.45◦ ), respec-
oxidation power of I-doped TiO2 may result in an increased CO2 tively. Because the brookite (1 2 0) (2 = 25.34◦ ) and brookite (1 1 1)
photoreduction rate. (2 = 25.69◦ ) peaks overlap with the anatase (1 0 1) peak, AA and
In this study, I-doped TiO2 photocatalysts were synthesized via AB were calculated by the following method. Using the single
a hydrothermal method and were evaluated for the first time for isolated brookite (1 2 1) peak as a reference, the anatase (1 0 1),
CO2 reduction under UV and visible light irradiation. The effects brookite (1 2 0) and brookite (1 1 1) overlapped peaks were decon-
of iodine doping levels and calcination temperature on the cat- (1 2 1) (1 2 0)
voluted by 0.9 and 0.8 intensity ratio for I(brookite) /I(brookite) and
alytic activity were also investigated, a topic that has been scarcely
(1 1 1) (1 2 0)
discussed in the literature for I-doped TiO2 . The structural and com- I(brookite) /I(brookite) respectively, with the same FWHM of brookite
positional properties of the I-TiO2 nanomaterials were analyzed (1 2 1) [35].
and correlated with their photocatalytic reduction performance. The lattice structure of individual nanocrystals was visualized by
This is the first paper that reports photocatalytic CO2 reduction phase-contrast high resolution transmission electron microscopy
by nonmetal doped TiO2 without any other co-catalysts. There- (HRTEM) carried out with 300 keV electrons in a Hitachi H9000NAR
fore, the findings are an important step towards the discovery of instrument with 0.18 nm point and 0.11 nm lattice resolution.
cost-effective catalysts for CO2 reduction to solar fuels. Two-dimensional Fourier transforms were calculated and used to
measure lattice spacing and interplanar angles. Amplitude con-
trast TEM images were used to obtain direct information about
2. Experimental the nanocrystal sizes. Selected area electron diffraction (SAD) pro-
vided information that is analogous to XRD, but from nanocrystals
2.1. Photocatalyst preparation supported on an electron-transparent amorphous carbon film and
selected within a ∼450 nm diameter aperture.
The method for synthesis of I-doped TiO2 was modified from The UV–vis diffuse reflectance spectra were obtained by a
that reported by Tojo et al. [33]. The preparation started by dissolv- UV–vis spectrometer (Ocean Optics) using BaSO4 as the back-
Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202 197

Fig. 1. Experimental setup for CO2 photoreduction. 1: Mass flow controller; 2: water
bubbler; 3: photoreactor with a quartz window; 4: two-way valve; 5: long-pass
filter; 6: gas chromatograph (GC/TCD-FID); 7: catalyst samples dispersed on glass-
fiber filter; 8: Xe lamp; 9: sampling port.

ground. The reflectance was converted to F(R) values using the


Kubelka–Munk function (Eq. (4)).

(1 − R)2 k
F(R) = = (4)
2R s
where R is the absolute reflectance, k is the molar absorption
coefficient and s is the scattering coefficient. The band gaps were
obtained from the plot of [F(R)Eph ]1/2 against the photon energy Fig. 2. XRD patterns of I-doped TiO2 at different calcination temperatures (a) and
Eph . X-ray photoelectron spectroscopy (XPS) analysis was carried I-doped TiO2 at different iodine doping level (b) (A: anatase; B: brookite; R: rutile).
out on a PerkinElmer PHI 5100 ESCA system with an Al K␣ X-ray
source (h = 1253.6 eV) and pass energy of 35.75 eV operating at
filter in the reactor. Again, for either case no carbon-containing
a pressure of 8 × 10−10 Torr. The observed spectra were corrected
compounds were produced under either UV or visible irradiation.
with the C1s binding energy (BE) value of 284.6 eV.
This demonstrates that the reactor and the glass-fiber filter were
clean and that the CO2 conversion cannot proceed without the
2.3. CO2 photoreduction experiments
photocatalyst. All these background tests have proved that any
carbon-containing compounds produced must be originated from
The schematic of the photocatalytic reaction system is illus-
CO2 through photocatalytic reactions.
trated in Fig. 1. Compressed CO2 (99.99%, Praxair) regulated by a
mass flow controller was passed through a water bubbler to gener-
ate CO2 and H2 O vapor mixture (H2 O, v/v% ≈ 2.3%). The gas mixture 3. Results and discussion
was then purged through a cylindrical photoreactor (V = 58 cm3 )
with stainless steel walls and a quartz window. A fixed amount of 3.1. Average nanocrystal structure from XRD analysis
powder catalyst (200 mg) was dispersed on a glass-fiber filter and
placed at the bottom of the reactor. After purging for 1.5 h, the gas Fig. 2 shows the XRD patterns of TiO2 samples doped with dif-
valves on both sides of the reactor were closed to seal the reactor. ferent concentrations of iodine calcined at 375 ◦ C and 5% I-TiO2
A 450 W Xe lamp (Oriel) was used as the light source and a long- calcined at different temperatures. The calculated values of phase
pass filter was applied to cut off the short wavelengths that are content and crystal size are listed in Table 1. The undoped and I-
less than 400 nm if only visible light is needed. A spectroradiome- doped TiO2 mainly consist of two phases, anatase and brookite.
ter (International Light Technologies ILT950) was used to obtain As the calcination temperature increased from 375 to 450 ◦ C, the
the spectral intensity of the Xe lamp with and without the filter. anatase phase content increased and the brookite phase decreased;
During the illumination period, the gaseous samples in the reactor when the calcination temperature increased to 550 ◦ C, the brookite
were taken by a gastight syringe (Hamilton, #1750, 500 ␮l) every content further decreased with the appearance of a small percent-
30 min and manually injected to a gas chromatograph (GC, Agilent age of rutile. The phase transition between metastable anatase and
7890A) equipped with both a thermal conductivity detector (TCD) brookite is not well studied in the literature; however, the result
and flame ionization detector (FID). in this study seems to be in agreement with some of the litera-
Prior to CO2 photoreduction experiments, all catalyst powders ture that upon calcination brookite transforms to rutile via anatase
were pre-treated under UV irradiation (12 W, 365 nm) for 12 h to [36–38]. In other words, brookite first transforms to anatase
eliminate organic residues on the catalyst, if any. GC measurements and then anatase transforms to rutile. In contrast to undoped
were also performed using a mixture of ultra high purity helium nanocrystal TiO2 whose anatase-to-rutile transformation temper-
(instead of CO2 ) and water vapor as the purging and reaction gas ature is around 700 ◦ C [39], the lower temperature (450–550 ◦ C)
for the catalyst-loaded reactor; no carbon-containing compounds of transformation to rutile in this study was possibly due to the
were produced by the catalyst under UV or visible irradiation. This dopant-induced instability of TiO2 caused by lattice distortion and
verifies that the catalyst was clean (i.e. no interference from organic bond weakening [33], even at a low dopant concentration (for 5%
residues). A series of other background tests were also conducted I-TiO2 ). When the calcination temperature was kept the same at
using a mixture of CO2 and H2 O vapor as the purging and reaction 375 ◦ C, increasing the iodine concentration only slightly decreased
gas for both cases of (1) empty reactor and (2) blank glass-fiber the brookite content by a few percent.
198 Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202

Table 1
Phase content and average crystal size of I-TiO2 samples obtained from X-ray diffraction, band gap from optical spectroscopy, and specific surface area from BET analysis
(A:anatase, B:brookite, R: rutile).

Sample Phase content (%) Crystal size (nm) Band gap (eV) BET specific surface area(m2 /g)

A B R A B

TiO2 -375C 66 34 0 8.8 4.7 3.13 122.9


5% I-TiO2 -375C 66 34 0 7.5 5.4 3.05 137.4
5% I-TiO2 -450C 71 29 0 8.4 10.1 – 99.4
5% I-TiO2 -550C 76 19 5 19.1 12.7 – 43.1
10% I-TiO2 -375C 70 30 0 5.5 6.3 3.00 137.6
15% I-TiO2 -375C 72 28 0 5.8 5.5 3.02 137.6

In terms of crystal size, it is clear from Table 1 that with increas-


ing calcination temperature the average crystal size increases for
both anatase and brookite crystals (by a factor of ∼2.45 ± 0.10 at
550 ◦ C compared with 375 ◦ C). At the same calcination tempera-
ture of 375 ◦ C, the undoped TiO2 has the largest crystal size for
anatase (8.8 nm), and the crystal size decreases (to 7.5, 5.5, and
5.8 nm) as the nominal iodine doping level increases in increments
of 5% (from 0% to 15%). This result agrees with the literature that
dopants can favor the formation of smaller particles [40]. For exam-
ple, Zhou et al. [22] reported the particle sizes of N doped and N–I
co-doped TiO2 are smaller than pure TiO2 . Su et al. [30] found that I-
doped TiO2 has much smaller crystallite size (7.7 nm, anatase) than
undoped TiO2 (23.7 nm, anatase) and suggested that the repulsion
among adsorbed iodine species inhibits crystal growth. An interest-
ing finding in our study is that the brookite crystallite size of I-doped
TiO2 is slightly larger than that of undoped one (Table 1), indicat-
ing that iodine has opposite effect on the growth rate of anatase
and brookite nanocrystals under otherwise identical hydrother-
mal conditions. The lack of literature on doping of brookite TiO2
warrants further investigation in this interesting phenomenon.

3.2. Individual nanocrystal structure and morphology from TEM


analysis

Fig. 3 shows amplitude-contrast transmission electron


microscopy (TEM in (a)) and phase-contrast high-resolution
TEM images (HRTEM in (b) and (c)) of the 5% I-TiO2 -375C sample.
Both types of images show agglomerates of TiO2 nanocrystals.
The crystallite size is in the range of 6 – 9 nm, which is in good
agreement with the average size calculated from the Scherrer
equation. Similarly, selected area electron diffraction experiments
(SAED inset in (a)) recorded from agglomerates within a selecting
aperture of 450 nm confirm the phase determination of XRD
and demonstrates that the anatase and brookite phases of I-TiO2
occur in close proximity. The HRTEM images show lattice fringes
within individual nanocrystals. Analysis of the lattice spacings and
interpanar angles finds that each nanocrystal has a well-defined
phase and the lattice appears cleanly and bulk-terminated at the
surface. The nanocrystal morphology is defined by low-energy
facets that are conjoined by curved surfaces composed of closely
spaced terraces and steps. For example, the HRTEM image in Fig. 3c
shows clear one-dimensional lattice fringes of TiO2 (lattice spac-
ing = 0.345 nm) which is very close to the brookite TiO2 (1 1 1) bulk
lattice spacing of 0.346 nm, according to powder diffraction file
(PDF) No. 29-1360. Since all of the anatase spacings overlap with
brookite very closely, it is only possible to uniquely determine the
termination facets of the brookite nanocrystals. These consistently
yield the (1 1 1) type of crystal plane as dominant facet for the
brookite TiO2 nanocrystals. The second type of termination plane
occurs for interplanar distance of ∼0.351 nm which is the (1 0 1) Fig. 3. Electron microscopy of 5% I-TiO2 -375C sample: (a) TEM image and SAED
plane of anatase or the (1 2 0) plane of brookite. It is unlikely that (inset), (b) HRTEM image with labeled examples of anatase (A) and brookite (B)
brookite nanocrystals would have two very different dominant nanocrystals, and (c) HRTEM lattice spacings and dominant surface facets for A (1 2 0)
and B (1 1 1) nanocrystals, with arrows pointing at steps on B (1 1 1) surface.
facets under the same growth and calcination condition. Hence,
Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202 199

and 43.1 m2 /g, respectively, which corresponds well to the increase


in crystal size.

3.5. Host and dopant valence from XPS studies

Fig. 5a shows XPS survey spectrum of 10% I-TiO2 calcined


at 375 ◦ C which indicates the existence of Ti, O, and I ele-
ments. Fig. 5b–d shows the high resolution XPS spectra scanning
over the following three binding energy areas: (1) Ti 2p region
(450–470 eV); (2) I 3d region (610–640 eV); and (3) O 1s region
(520–540 eV). Table 2 summarizes the surface atomic concentra-
tions of Ti, I, and O as well as the atomic percentages of the three
elements at different chemical states.
As shown in Fig. 5b, the XPS peaks of the 10% I-TiO2 -375C sam-
ple in the Ti 2p region appear at 458.5 eV(Ti 2p3/2 ) and 464.2 eV(Ti
2p1/2 ), both of which correspond to Ti4+ . There are two small peaks
in the lower side of Ti 2p3/2 (457.5 eV) and Ti 2p1/2 (463.0 eV), which
are ascribed to Ti3+ that is generated to maintain the electroneutral-
ity by I5+ substituting Ti4+ , as the ionic radii of I5+ and Ti4+ are very
close [32,41]. The XPS spectra of I 3d region (Fig. 5c) for 10% I-TiO2 -
375C show double peaks around 623.6 eV (I 3d5/2 ) and 635.1 eV (I
3d3/2 ) which infer that the oxidation state of doped iodine is I5+
[32,33]. Also two weaker satellite peaks around 619.9 eV (I 3d5/2 )
and 631.4 eV (I 3d3/2 ) indicate existence of I− [33,41]. The results
agree with the literature that I5+ /I− pairs were observed for I-doped
TiO2 and that I5+ ions substitute for Ti4+ and are present in the
I–O–Ti bond [28,33,41]. Some other studies have reported TiO2
doped with multivalency iodine I7+ /I− prepared by different meth-
ods or using different precursors [22,30]. The I7+ peak at around
Fig. 4. UV–vis diffuse reflectance spectra of TiO2 with different iodine doping levels 624.0 eV (I 3d5/2 ) was not observed in our study. The XPS spectra
(a) and plots of the square root of the Kubelka–Munk function versus the photon
energy (b).
of the O 1s region (Fig. 5d) show a major peak at around 529.5 eV
that that correspond to lattice oxygen O2− . The other peak around
530.8 eV can be attributed to chemisorbed oxygen on the surface
[22] or surface hydroxyl groups [33].
it is possible to conclude, by elimination, that the second type of From Table 2, it is clear that the surface atomic concentration of
facets belong to anatase (1 0 1) type planes. These morphology iodine increased with the nominal iodine doping level from 0% to
changes are the subject of on-going work and are beyond the scope 15%. The I/Ti atomic ratio on the surface is much larger than that
of this initial work to evaluate the efficacy of I-doped TiO2 as an in the bulk (nominal) for the 10% and 15% I-TiO2 samples, indicat-
effective material for CO2 photoreduction. ing that iodine is mainly doped on the surface. The percentage of
Ti3+ /Ti increased to 9.9%, 18.4% and 19.8% as the nominal iodine con-
3.3. Band gap analysis from UV–vis diffuse reflectance centration increased to 5%, 10% and 15% by weight. This supports
spectroscopy our conclusion that I5+ substitutes Ti4+ and results in generation of
Ti3+ due to charge imbalance. For the 15% I-TiO2 sample, the sur-
The UV–vis diffuse reflectance spectra of undoped TiO2 and I- face iodine concentration (2.5 at.%) and Ti3+ fraction (19.8 at.%) are
doped TiO2 samples are shown in Fig. 4a. The absorption edge of only slightly larger than those of the 10% I-TiO2 sample (2.3 at.%
undoped TiO2 is around 400 nm and is extended to the visible light and 18.4 at.%, respectively), suggesting that the doping of iodine
region for I-doped TiO2 with an iodine concentration from 2.5 to approaches to a saturation level for the 15% I-TiO2 sample. This also
15%, which matches the yellow color of the I-doped TiO2 . Fig. 4b correlates with the result that the band gap of I-TiO2 decreases with
illustrates the plots for obtaining the band gap values that are also increased iodine concentration but levels off at the concentration
listed in Table 1. The undoped TiO2 has a band gap of 3.13 eV, while of iodine greater than 10 wt.% (Table 1).
the band gaps of I-TiO2 slightly decrease with iodine doping and
level off at around 3.00 eV when the nominal iodine concentration 3.6. Photocatalytic activity for CO2 reduction
is greater than 10%.
In this study, CO was identified as the main CO2 reduction prod-
3.4. BET specific surface area uct using undoped and I-doped TiO2 , while our previous study
[13] showed that CH4 was produced in addition to CO when the
The BET specific surface areas (SSA) of the various I-TiO2 cata- TiO2 surface was loaded with Cu species. The following reactions
lysts are listed in Table 1. The SSA for undoped TiO2 is 122.9 m2 /g, may express the pathways of CO2 photoreduction to CO and water
which is much higher than that of commercially available P25 oxidation to O2 :
(∼50 m2 /g). With iodine doping in TiO2 , the SSA slightly increases
hv
as the crystal size slightly decreases in average, and all the I-TiO2 TiO2 −→e−
cb
+ h+
vb (R1)
samples calcined at the same temperature (375 ◦ C) have similar SSA
(∼137 m2 /g) since their average crystal sizes are very close to each 2H2 O + 4h+ → 4H+ + O2 (R2)
other, as seen in Table 1. Increasing the calcination temperature of
5% I-TiO2 to 450 and 550 ◦ C dramatically reduces the SSA to 99.4 CO2 + 2H+ + 2e− → CO + H2 O (R3)
200 Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202

Fig. 5. XPS spectra of 10% I-TiO2 calcined at 375 ◦ C: survey spectrum (a) and high-resolution spectra for Ti 2p (b), I 3d (c), and O 1s (d).

Fig. 6 shows the concentration of CO produced in the reactor (in less discussed possibly because they are more difficult to control.
ppm) as a function of illumination time under visible ( > 400 nm) This is the first study that has reported an optimal doping level for
and UV–vis irradiation ( > 250 nm), respectively. Undoped TiO2 I-doped TiO2 .
had no activity under visible light (results not shown in Fig. 6a). All The concentration of CO reached 670 ppm at 210 min for the
I-doped TiO2 showed visible light activity for CO2 photoreduction sample of 10% I-TiO2 -375C under visible light (Fig. 6a), resulting in
to CO and the concentration of CO increased almost linearly with a product yield equivalent to 2.4 ␮mol g−1 h−1 . There have been
illumination time (Fig. 6a). With the same iodine concentration very few studies on the CO2 photoreduction under visible light
(5%), I-TiO2 calcined at 375 ◦ C had the highest activity; increased irradiation. This paper, for the first time in the literature, reports
calcination temperatures (450 ◦ C and 550 ◦ C) lowered the CO2 pho- photocatalytic CO2 reduction by nonmetal-doped TiO2 without
toreduction rate. This is likely due to the significant increase of the any other co-catalysts. Grimes and co-workers [43] synthesized
crystal size and decrease in surface area with increasing calcination N-doped TiO2 nanotube arrays sputtered with Cu nanoparticles
temperature (Table 1). For I-TiO2 with the same calcination tem- (NT/Cu) for CO2 photoreduction under sunlight and they reported
perature (375 ◦ C), activity of CO2 photoreduction follows the order that the activity in visible light region is only 3% of that under
of 10% > 15% > 5% in terms of iodine doping concentration (Fig. 6a). the whole solar spectrum (the rates of CO, CH4 , and other HCs
The band gap analysis (Table 1) and XPS analysis data (Table 2) show production are equivalent to 0.11, 0.13, and 0.05 ␮mol g−1 h−1 ,
that 10% and 15% I-TiO2 samples have very close band gap energies respectively; the total production rate is about 0.3 ␮mol g−1 h−1
(3.00 eV and 3.02 eV, respectively) and surface iodine concentra- under visible light irradiation of the solar spectrum with an inten-
tions (2.3 at.% and 2.5 at.%, respectively), suggesting that the 15% sity of 78.5 mW/cm2 ) [43]. Ozcan et al. [44] studied dye-sensitized
I-TiO2 sample may not be superior to 10% I-TiO2 . Furthermore, too and Pt modified TiO2 for CO2 photoreduction and reported a CH4
high a dopant concentration may form charge recombination cen- production rate of 0.2 ␮mol g−1 h−1 using a 75 W daylight lamp as
ters and/or shield the surface of TiO2 from light irradiation, both of the visible light source. Fig. 7 shows the spectra of the 450 W Xe
which reduce the photocatalytic activity. These may explain why lamp used in this work, with or without the 400 nm long-pass fil-
10% corresponds to the optimal iodine concentration under visible ter, in comparison with the AM 1.5G standard solar spectrum. The
light irradiation. Similar findings on the optimal dopant concen- integrated light intensity of the Xe lamp was 428 mW/cm2 (full
tration have been reported for metal-doped TiO2 [3,13,17,41,42], spectrum) and 233 mW/cm2 for the visible region (400–750 nm).
while optimal doping concentrations of nonmetals have been much While the visible light intensity in our study was approximately

Table 2
Surface atomic concentration of I-TiO2 catalysts from XPS analysis.

Sample Surface atomic concentration (%) I/Ti atomic ratio (%) Ti species (%) I species (%) O species (%)
4+ 3+ 5+ −
I Ti O Nominal Surface Ti Ti I I Olattice Oabsorb

TiO2 -375C 0 17.1 82.9 – – 100 0 – – 65.3 34.7


5% I-TiO2 -375C 0.4 17.4 82.2 3.3 2.3 90.1 9.9 16.0 84.0 57.6 42.4
10% I-TiO2 -375C 2.3 20.2 77.5 7.0 11.4 81.6 18.4 69.9 30.1 56.3 43.7
15% I-TiO2 -375C 2.5 15.9 81.6 11.1 15.7 80.2 19.8 57.3 42.7 62.2 37.8
Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202 201

Fig. 8. Time dependence on the volumetric ratio of O2 /(O2 + N2 ) in the batch reactor
using 5% I-TiO2 375C as the photocatalyst.

for undoped TiO2 calcined at 375 ◦ C, confirming that undoped


TiO2 can only be activated by UV irradiation. However, the CO
production rate is very low for the pure TiO2 , especially when com-
pared with the I-doped TiO2 calcined at the same temperature. The
enhanced activity of I-doped TiO2 under UV–vis irradiation is likely
due to the combinational effects of slightly increased surface area,
increased visible light absorption, and improved charge separation
due to the iodine doping [33]. Among the three I-doped TiO2 (2.5%,
5%, and 10%), 5% I-TiO2 exhibited the highest activity followed by
10% I-TiO2 , and the CO production for 5% I-TiO2 reached nearly
600 ppm in 90 min and leveled off thereafter. The 15% I-TiO2 was
Fig. 6. Concentration of CO produced from CO2 photoreduction under visible light
not tested under UV–vis irradiation as its activity was inferior to
(a) and under UV–vis irradiation (b) using different I-doped TiO2 samples.
10% I-TiO2 under visible light irradiation.
It is interesting to find that the optimal iodine concentrations in
four times higher than those used by Grimes and co-workers [43] TiO2 are different under visible light (10%) and UV–vis (5%) irradi-
and Ozcan et al. [44], our CO2 photoreduction rate under visible ations, respectively. Furthermore, the highest CO production rate
light (2.4 ␮mol g−1 h−1 ) is approximately eight times higher than under visible light irradiation (670 ppm at 210 min for 10% I-TiO2 -
that reported by Grimes and co-workers [43] and twelve times 375C) is even higher than that under UV–vis irradiation (600 ppm
higher than that reported by Ozcan et al. [44]. Although the pho- at 90 min for 5% I-TiO2 -375C). In addition, the activity of 10% I-TiO2 -
toactivity may not be a linear function of light intensity, and the 375C under visible light irradiation is higher than that under UV–vis
product selectivities were different in the three studies, the I- irradiation, but the trend is opposite for 5% I-TiO2 -375C. These
TiO2 photocatalyst developed in this work is a very attractive and results indicate that materials designed for high activity under vis-
promising candidate for larger scale CO2 photoreduction under ible light may not be necessarily optimized for UV applications.
sunlight. This catalyst is highly responsive to visible light and its Similar findings have been reported that N- or S-doped TiO2 has
synthesis requires only low-cost materials and simple manufac- superior photooxidation activity to that of undoped TiO2 under vis-
turing process steps. ible light irradiation but has similar or even lower photocatalytic
Fig. 6b shows the CO2 photoreduction activity of the catalysts activity in the UV region [45,46].
under UV–vis irradiation. CO2 photoreduction to CO was observed To verify the half reaction of O2 production from H2 O oxidation,
the concentration of O2 was monitored in a separate test using 5%
I-TiO2 375C as the catalyst under UV–vis irradiation. There were
background O2 and N2 (a few hundred ppm) detected in the reactor,
possibly because the reactor was not vacuumed before purging it
with the CO2 –H2 O mixture. In addition, the GC peaks for O2 and N2
were close to each other and overlapped to a certain degree. Hence,
the change of volumetric ratio of O2 /(O2 + N2 ) in the batch reactor
would be a better indicator for O2 production during the photo-
catalytic reaction. As shown in Fig. 8, the O2 /(O2 + N2 ) ratio slightly
decreased in the first 30 min and then increased with the irradiation
time at an almost linear rate, implying O2 production from H2 O dis-
sociation according to R2. The increasing CO concentration under
light irradiation (Fig. 6) together with the increasing O2 /(O2 + N2 )
ratio (Fig. 8) provide sound evidence of photocatalytic reaction of
CO2 with H2 O to form CO and O2 . The possibility of H2 production
due to photocatalytic H2 O reduction was also tested by switching
the GC carrier gas from helium to nitrogen to enhance the sensitiv-
ity for H2 . However, under identical experimental conditions using
Fig. 7. Spectra of the Xe lamp used in this study, with and without the 400 nm the 5% I-TiO2 375C, no H2 was detected during a three hour UV–vis
long-pass filter, in comparison to the AM 1.5G standard solar spectrum. illumination. H2 O reduction to produce H2 , which competes with
202 Q. Zhang et al. / Applied Catalysis A: General 400 (2011) 195–202

CO2 reduction for electrons, is feasible based on the thermochem- [7] K. Teramura, T. Tanaka, H. Ishikawa, Y. Kohno, T. Funabiki, J. Phys. Chem. B 108
ical data but experimental results from this study and from the (2004) 346–354.
[8] Z.Y. Wang, H.C. Chou, J.C.S. Wu, D.P. Tsai, G. Mul, Appl. Catal. A: Gen. 380 (2010)
literature [18,47] suggest that the yield of H2 is trace, if any, as com- 172–177.
pared with those of the CO2 reduction products, and is prominent [9] K. Teramura, H. Tsuneoka, T. Shishido, T. Tanaka, Chem. Phys. Lett. 467 (2008)
only in the presence of noble metals (e.g. Ru, Pt). 191–194.
[10] T. Hirose, Y. Maeno, Y. Himeda, J. Mol. Catal. A: Chem. 193 (2003) 27–32.
[11] K. Koci, L. Obalva, Appl. Catal. B: Environ. 89 (2009) 494–502.
4. Conclusions [12] M. Anpo, H. Yamashita, K. Ikeue, Y. Fujii, S.G. Zhang, Y. Ichihashi, D.R. Park, Y.
Suzuki, K. Koyano, T. Tatsumi, Catal. Today 44 (1998) 327–332.
[13] Y. Li, W.-N. Wang, Z. Zhan, M.-H. Woo, C.-Y. Wu, P. Biswas, Appl. Catal. B:
I-doped TiO2 photocatalysts with different iodine doping levels Environ. 100 (2010) 386–392.
and calcination temperatures were prepared and for the first time [14] K. Koci, L. Obalova, Z. Lacny, Chem. Pap. 62 (2008) 1–9.
were tested for the activity of photocatalytic CO2 reduction with [15] S. Rengaraj, X.Z. Li, J. Mol. Catal. A: Chem. 243 (2006) 60–67.
[16] N.L. Wu, M.S. Lee, Int. J. Hydrogen Energy 29 (2004) 1601–1605.
H2 O. The iodine dopant extended the absorption spectra of TiO2 to [17] K. Koci, K. Mateju, L. Obalova, S. Krejcikova, Z. Lacny, D. Placha, L. Capek, A.
the visible light region and facilitated charge separation. XPS anal- Hospodkova, O. Solcova, Appl. Catal. B: Environ. 96 (2010) 239–244.
ysis revealed that I5+ substitutes Ti4+ in the lattice and as a result, [18] N. Sasirekha, S.J.S. Basha, K. Shanthi, Appl. Catal. B: Environ. 62 (2006) 169–180.
[19] R. Dholam, N. Patel, Int. J. Hydrogen Energy 34 (2009) 5337–5346.
Ti3+ is generated to balance the charge. Significant enhancement of
[20] M. Hamadanian, A. Reisi-Vanani, A. Majedi, Mater. Chem. Phys. 116 (2009)
CO2 photoreduction to CO was observed for I-doped TiO2 compared 376–382.
with undoped TiO2 under both visible and UV–vis irradiations. [21] C. Wen, Y.J. Zhu, T. Kanbara, H.Z. Zhu, C.F. Xiao, Desalination 249 (2009)
The high activity of CO2 reduction under visible light compared to 621–625.
[22] L. Zhou, J. Deng, Y. Zhao, W. Liu, L. An, F. Chen, Mater. Chem. Phys. 117 (2009)
the literature data makes I-doped TiO2 a potentially cost-effective 522–527.
photocatlayst. Lower calcination temperature (375 ◦ C) resulted in [23] S. In, A. Orlov, R. Berg, F. Garcia, S. Pedrosa-Jimenez, M.S. Tikhov, D.S. Wright,
smaller particle size and higher catalytic activity. 10% I-doping level R.M. Lambert, J. Am. Chem. Soc. 129 (2007) 13790–13791.
[24] S.U.M. Khan, M. Al-Shahry, W.B. Ingler, Science 297 (2002) 2243–2245.
demonstrated the highest CO2 photoreduction rate under visible [25] J.C. Yu, J.G. Yu, W.K. Ho, Z.T. Jiang, L.Z. Zhang, Chem. Mater. 14 (2002)
light, while 5% I-doping level performed best under UV–vis irradi- 3808–3816.
ation. Too high an iodine doping level may result in recombination [26] G.S. Wu, J.L. Wen, J.P. Wang, D.F. Thomas, A.C. Chen, Mater. Lett. 64 (2010)
1728–1731.
centers and thus lower the photocatalytic activity. Further work is [27] M. Pelaez, A.A. de la Cruz, E. Stathatos, P. Falaras, D.D. Dionysiou, Catal. Today
needed to determine if the nanocrystal morphology changes as a 144 (2009) 19–25.
function of doping level and annealing conditions, especially for the [28] X.T. Hong, Z.P. Wang, W.M. Cai, F. Lu, J. Zhang, Y.Z. Yang, N. Ma, Y.J. Liu, Chem.
Mater. 17 (2005) 1548–1552.
preparations that show superior photocatalytic activity. [29] G. Liu, C.H. Sun, X.X. Yan, L. Cheng, Z.G. Chen, X.W. Wang, L.Z. Wang, S.C. Smith,
G.Q. Lu, H.M. Cheng, J. Mater. Chem. 19 (2009) 2822–2829.
Acknowledgements [30] W.Y. Su, Y.F. Zhang, Z.H. Li, L. Wu, X.X. Wang, J.Q. Li, X.Z. Fu, Langmuir 24 (2008)
3422–3428.
[31] R. Long, Y. Dai, B.B. Huang, Comp. Mater. Sci. 45 (2009) 223–228.
This work was supported by start-up funds from the Uni- [32] Z.Q. He, X. Xu, S. Song, L. Xie, J.J. Tu, J.M. Chen, B. Yan, J. Phys. Chem. C 112 (2008)
versity of Wisconsin-Milwaukee (UWM) and the UWM Research 16431–16437.
[33] S. Tojo, T. Tachikawa, M. Fujitsuka, T. Majima, J. Phys. Chem. C 112 (2008)
Foundation Bradley Catalyst Grant (Li), and by Department
14948–14954.
of Energy DE-FG02-06ER46329 (Gajdardziska-Josifovska). The [34] H.Z. Zhang, J.F. Banfield, J. Phys. Chem. B 104 (2000) 3481–3487.
authors would like to acknowledge technical assistance by Steven [35] S. Kaewgun, C.A. Nolph, B.I. Lee, L.Q. Wang, Mater. Chem. Phys. 114 (2009)
Hardcastle and Donald Robertson from the Advanced Analysis 439–445.
[36] A. Amorelli, J.C. Evans, C.C. Rowlands, J. Chem. Soc. Faraday Trans. 85 (1989)
Facility and the Physics Laboratory for High Resolution Transmis- 4031–4038.
sion Electron Microscopy at UWM. They would also like to thank [37] S. Bakardjieva, V. Stengl, L. Szatmary, J. Subrt, J. Lukac, N. Murafa, D. Niznansky,
Dr. Chang-Yu Wu at University of Florida for assistance in XPS and K. Cizek, J. Jirkovsky, N. Petrova, J. Mater. Chem. 16 (2006) 1709–1716.
[38] K.R. Zhu, M.S. Zhang, J.M. Hong, Z. Yin, Mater. Sci. Eng. A: Struct. Mater. Prop.
BET analysis. Microstruct. Process. 403 (2005) 87–93.
[39] W. Li, C. Ni, H. Lin, C.P. Huang, S.I. Shah, J. Appl. Phys. 96 (2004) 6663–6668.
References [40] X.L. Nie, S.P. Zhuo, G. Maeng, K. Sohlberg, Int. J. Photoenergy 2009 (2009) 1–22.
[41] K.X. Song, J.H. Zhou, J.C. Bao, Y.Y. Feng, J. Am. Ceram. Soc. 91 (2008) 1369–1371.
[42] H. Yamashita, H. Nishiguchi, N. Kamada, M. Anpo, Y. Teraoka, H. Hatano, S.
[1] C.M. White, B.R. Strazisar, E.J. Granite, J.S. Hoffman, H.W. Pennline, J. Air Waste Ehara, K. Kikui, L. Palmisano, A. Sclafani, M. Schiavello, M.A. Fox, Res. Chem.
Manage. Assoc. 53 (2003) 645–715. Intermed. 20 (1994) 815–823.
[2] M. Anpo, H. Yamashita, Y. Ichihashi, S. Ehara, J. Electroanal. Chem. 396 (1995) [43] O.K. Varghese, M. Paulose, T.J. LaTempa, C.A. Grimes, Nano Lett. 9 (2009)
21–26. 731–737.
[3] I.H. Tseng, W.C. Chang, J.C.S. Wu, Appl. Catal. B: Environ. 37 (2002) 37–48. [44] O. Ozcan, F. Yukruk, E.U. Akkaya, D. Uner, Top. Catal. 44 (2007) 523–528.
[4] J.C.S. Wu, H.M. Lin, C.L. Lai, Appl. Catal. A: Gen. 296 (2005) 194–200. [45] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269–271.
[5] M.C. Hidalgo, G. Colon, J.A. Navio, A. Macias, V. Kriventsov, D.I. Kochubey, M.V. [46] T. Ohno, T. Mitsui, M. Matsumura, Chem. Lett. 32 (2003) 364–365.
Tsodikov, Catal. Today 128 (2007) 245–250. [47] K. Koci, L. Obalova, L. Matejova, D. Placha, Z. Lacny, J. Jirkovsky, O. Solcova, Appl.
[6] C.C. Lo, C.H. Hung, C.S. Yuan, J.F. Wu, Solar Energy Mater. Solar Cells 91 (2007) Catal. B: Environ. 89 (2009) 494–502.
1765–1774.

You might also like