You are on page 1of 8

Supplementary Information for the Journal of American Chemical Society

© The American Chemical Society 2014

Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by


Coexposed {001} and {101} Facets
† † † †
Jiaguo Yu,*, Jingxiang Low, Wei Xiao, Peng Zhou, Mietek Jaroniec*,‡

†State Key Laboratory of Advanced Technology for Material Synthesis and Processing, Wuhan
University of Technology, Luoshi Road 122#, Wuhan 430070, P. R. China.
‡ Department of Chemistry and Biochemistry, Kent State UniVersity, Kent, Ohio 44242, USA.
Email: jiaguoyu@yahoo.com; jaroniec@kent.edu

Experimental Section

Computational details. The DOS plots of anatase TiO2 were obtained by using the first-
principles DFT calculations to specify the electronic structure of {001} and {101} facets in this
photocatalyst. CASTEP package based on the plane-wave-pseudo potential approach was used to
perform DFT calculations.

Sample preparation. TiO2 samples were synthesized by using a solvothermal method. In this
synthesis 25 mL of tetrabutyl titanate were mixed with 0, 3, 4.5, 6 and 9 mL of HF solution (with
a concentration 40 wt%), respectively, under stirring for 30 min. Then, the solution was
transferred into 100 mL Teflon-lined autoclave and kept at 180 ◦C for 24 h. After solvothermal
reaction, the resulting white precipitates were collected and washed with ethanol and distilled
water for three times. Then, the samples were dried in an oven at 80 ◦C for 6 h. The prepared
samples were denoted as HF0, HF3, HF4.5, HF6, and HF9, respectively, according to the
amount of hydrofluoric acid solution added. To remove the surface fluorine ions from the
prepared samples, all the as-prepared samples were heated at 550 ◦C for 2 h in a furnace. Then,
the samples were cooled to room temperature.

Photocatalytic experiments. The photocatalytic reduction of CO2 was carried out in a 200
mL home-made Pyrex reactor at ambient temperature and atmospheric pressure. A 300 W
simulated solar Xe arc lamp was used as the light source and positioned 10 cm above the
photocatalytic reactor. In a typical photocatalytic reduction experiment, 100 mg of the sample
was put into the glass reactor and added to 10 mL of deionized water. The catalyst was
ultrasonically dispersed for 30 min, then, dried at 80 oC for 2 h to evaporate water. The sample
deposited on the bottom of the reactor was in the form of a thin film. Before the irradiation, the
reactor was sealed and blown with nitrogen for 30 min to remove air and ensure that the reaction
system was under anaerobic conditions. CO2 and H2O vapour were in-situ generated by the
reaction of NaHCO3 (0.12 g; introduced into the reactor before sealing) and HCl aqueous
S1
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

solution (0.25 mL, 4M), which was introduced into the reactor by syringe. 1 mL of mixed gas
was taken out from the reactor at given intervals (1 h) during the irradiation and analyzed by
using a gas chromatograph (GC-2014C, Shimadzu, Japan) equipped with a flame ionized
detector (FID) and methanizer. Products were analyzed on the basis of retention time data for the
analyte and the standard gaseous sample. The carrier gas used in the GC-2014C was high purity
nitrogen. The reaction product was also analyzed by another gas chromatograph (GC-14C,
Shimadzu, Japan, TCD, 5 Å molecular sieve column) with nitrogen used as a carrier gas. Blank
experiments were carried out in the absence of CO2 or light irradiation to confirm that CO2 and
light were two key factors influencing the photocatalytic CO2 reduction.

Sample characterization. Morphological observations were carried out on an S-4800 field


emission scanning electron microscope (SEM, Hitachi, Japan) and connected with an Oxford
Instruments X-ray analysis system. TEM analyses were performed with a JEM-2100 electron
microscope (JEOL, Japan), using a 200 kV accelerating voltage. The XRD was conducted on an
X-ray diffractometer (type HZG41B-PC) using Cu K radiation at a scan rate of 0.05◦2s−1. The
accelerating voltage and applied current were 40 kV and 80 mA, respectively. The SBET specific
surface area of the powders was determined by using nitrogen adsorption data recorded on a
Micromeritics ASAP 2020 nitrogen adsorption apparatus (USA). All the samples were degassed
at 180 ◦C prior to nitrogen adsorption measurements. The BET surface area was obtained by a
multipoint BET method using the adsorption data in the relative pressure (P/P0) range of 0.05–
0.3. The desorption branch of the isotherm was used to obtain the pore size distribution via the
Barret–Joyner–Halender (BJH) method derived for cylindrical pores. The volume adsorbed at
the relative pressure (P/P0) of 0.994 was used to determine the pore volume and the average pore
size. The X-ray photoelectron spectroscopy (XPS) measurement was obtained in an ultrahigh
vacuum VGESCALAB 210 electron spectrometer equipped with a multichannel detector, using
Mg K (1253.6 eV) radiation (operated at 200 W) of a twin anode in the constant analyzer
energy mode with a pass energy of 30 eV. All the binding energies were referenced to the C 1s
peak at 284.8 eV of the surface adventitious carbon.

Electrochemical test. EIS was tested at the open circuit potential. Electrochemical analyzer
(CHI660C instruments, CHI, China) in a standard three-electrode system was used to perform
the electrochemical measurement. The photocatalyst sample was coated onto a 2 cm × 1.5 cm
fluorine-tin oxide (FTO) glass electrode, and then annealed at 400 oC for 1 h. Na2SO4 aqueous
solution (1 M) was used as the electrolyte in this system. The sample coated FTO, Pt wire and
Ag/AgCl were used as the working electrode, counter electrode, and reference electrode in the
electrochemical analyzer system, respectively. A low power UV-LEDs (3 W, 365 nm) was used
as a light source.

S2
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

Figure S1. (a) The thermodynamically stable anatase TiO2 under equilibrium conditions and (b)
the growth of anatase TiO2 along [001] and [101] directions according to the Wulff construction.

Figure S2. Schematic of atomic surface structure of anatase TiO2 (a) {101} and (b) {001} facets.

S3
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

(A) (B)

a 001 a I
F
b 101 b θ H
G

Figure S3. A) Slab model of anatase TiO2 single crystal. B) Equilibrium model of anatase TiO2
single crystal.
C) Percentage of {001} facets can be calculated as follows [1-4]:
S001  2a 2
S101  8( 12 EG  b  12 EF  a)
S 001
S 001 % 
S 001  S101

2a 2

2a 2  8( 12 EG  b  12 EF  a)

a2
 1
b 1
a
a 2  4( 12  2
 b  12 2  a)
cos  cos 
a2 1
 
b2  a 2
b2
1
a2  1 a2

cos  cos 
cos

b2
cos  2  1
a

S4
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

cos

a
cos  ( )  2  1
b
Here θ is the theoretical value (68.3o) for the angle between the [001] and [101] facets of anatase.
As indicated in the slab model, two independent parameters b and a denote lengths of the side of
the bipyramid and the side of the square {001} ‘truncation’ facets, respectively.
References
[1] Xiang, Q. J.; Yu, J. G.; Jaroniec, M. Chem. Commun., 2011, 47, 4532.
[2] Yang, H. G.; Sun, C. H.; Qiao, S. Z.; Zou, J.; Liu, G.; Smith, S. C.; Cheng, H. M.; Lu, G. Q.
Nature, 2008, 453, 638.
[3] Zhang, D. Q.; Li, G. S.; Yang, X. F.; Yu, J. C. Chem. Commun., 2009, 4381.
[4] Yu, J. G.; Dai, G. P.; Xiang, Q. J.; Jaroniec, M. J. Mater. Chem., 2011, 21, 1049.
(101)
Relative intesity (a.u.)

(200)
(004)

(211)
(105)

HF9 (204)
HF6
HF4.5
HF3
HF0

10 20 30 40 50 60 70
2 Theta (degree)

Figure S4. XRD patterns of the TiO2 samples prepared by varying the HF amount.

S5
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

O KLL
O 1s
Ti 2p
Relative intensity (a.u.)

Ti 2s
C 1s
Ti 3p
HF 4.5
Ti 3s

HF 0

0 200 400 600 800


Binding energy (eV)
Figure S5. XPS survey spectra of HF0 and HF4.5.

HF0
Absorbance (a.u.)

HF4.5
HF9

0
350 400 450 500 550 600
Wavelength (nm)

Figure S6. UV-Vis absorption spectra of HF0, HF4.5 and HF9.

S6
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

300 HF0

Volume adsorbed (cm /g,STP)


HF4.5
250

3
200
1 10 100
150 Pore diameter (nm)

100

50

0
0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (P/P0)
Figure S7. Nitrogen adsorption-desorption isotherms and the corresponding pore size
distribution curves (inset) of HF0 and HF4.5.

Figure S8. EIS spectra of the TiO2 samples prepared by varying the HF amount.

S7
Supplementary Information for the Journal of American Chemical Society
© The American Chemical Society 2014

Table S1. Comparison of the specific surface areas, activity and specific activity of the HF0,
HF4.5 and HF9 samples.

Activity Specific
BET PV APS
Sample activity
molh-1g-1)
(m2/g) (cm3/g) (nm)
molh-1m-2)

HF0 107 0.32 11.9 0.15 1.4 x 10-3

HF4.5 45 0.22 15.1 1.35 30 x 10-3

HF9 30 0.12 16.8 0.55 18.3 x 10-3


PV: Pore volume; APS: Average pore size

S8

You might also like