Вы находитесь на странице: 1из 25

Subscriber access provided by Nanyang Technological Univ

Article
A New Method to Synthesize S-Doped TiO2 with Highly Efficient
and Stable Indoor Sunlight Photocatalytic Performance
Mingshan Zhu, Chunyang Zhai, Liqun Qun, Cheng Lu, Andrew S Paton, Yukou Du, and M. Cynthia Goh
ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/
acssuschemeng.5b01137 • Publication Date (Web): 02 Nov 2015
Downloaded from http://pubs.acs.org on November 12, 2015

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the
course of their duties.
Page 1 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4
5
6
7
8 A New Method to Synthesize S-Doped TiO2
9
10
11
with Stable and Highly Efficient
12
13 Photocatalytic Performance under Indoor
14
15
16 Sunlight Irradiation
17
18
19
20
Mingshan Zhu †, Chunyang Zhai ‡, Liqun Qiu †, Cheng Lu †,*, Andrew S. Paton †,
21
22 Yukou Du ‡, and M. Cynthia Goh †,*
23
24
25 † Department of Chemistry, University of Toronto, M5S3H6, Canada.
26
27
28 ‡ College of Chemistry, Chemical Engineering and Materials Science, Soochow
29
30 University, Suzhou 215123, China.
31
32
33 ABSTRACT In this paper, we report a new, low-cost and facile solvothermal
34
35
36 approach to synthesize visible-light-active S-doped TiO2 (S-TiO2) by using dimethyl
37
38 sulfoxide (DMSO) as both the S source and the solvent. Energy-dispersive X-ray
39
40
41 (EDX) and X-ray photoelectron spectroscopy (XPS) solidly confirmed the presence of
42
43
44 S element in the final product. The as-prepared S-TiO2 nanoparticles exhibited
45
46 excellent and long-term stable photocatalytic performance for the degradation of
47
48
49 organic pollutants under visible and indoor sunlight illumination. The catalyst still
50
51
52 kept high photoactivity even after several months of exposure to the indoor sunlight
53
54 irradiation. This result suggests a new approach to achieve stable and highly efficient
55
56
57 solar light driven photocatalysts for water purification.
58
59
60
KEYWORDS: photocatalyst; S-doped TiO2; visible light; indoor sunlight; long-term

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 2 of 24

1
2
3
4 stability
5
6
7
8 INTRODUCTION
9
10
11 To cope with the growing pollution of our hydrosphere, a variety of technologies,
12
13 including heterogeneous semiconductor photocatalytic oxidation, have been
14
15
16 developed for wastewater treatment.1-4 Titanium dioxide (TiO2), as the most widely
17
18
19 studied semiconductor photocatalyst, has thus far been explored to meet the
20
21 requirements of water purification.1-4 However, the band gap of pure TiO2 is ca. 3.2
22
23
24 eV, which means that it can only show activity under UV irradiation. It is well known
25
26
27 that UV light accounts for no more than 5% of the total solar energy, which is a small
28
29 amount compared to the 45% of energy in the visible region.4 Hence, in order to
30
31
32 effectively utilize solar radiation, it is desirable to develop efficient
33
34
35 visible-light-driven photocatalysts for remediating the growing pollution in our
36
37 hydrosphere.
38
39
40 Since the Asahi group first reported the visible-light photocatalytic activity of
41
42
43 nitrogen doped TiO2,5 more research has been focused on the modification of TiO2
44
45 with non-metal or/and metal ion doping.3-5 In these cases, the dopant is often
46
47
48 incorporated as an anion or cation to take the place of Ti or/and oxygen in the lattice
49
50
51 of TiO2, resulting in bandgap narrowing in TiO2 nanostructures and showing high
52
53 visible-light photocatalytic activity. Among these dopants, sulfur (S) doping has
54
55
56 received particular attention, owing to its highly thermal stability and significant
57
58
59 enhancement in visible light driven photocatalytic activity,6-13 where the TiS2, thiourea,
60
CS2, etc. were often used as the S source.6-13 However, these precursors are either
2

ACS Paragon Plus Environment


Page 3 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 expensive or highly toxic.
5
6
7 Dimethyl sulfoxide (DMSO), an important polar aprotic compound, is widely
8
9
10 used as a solvent in various organic and inorganic syntheses. Recently, some
11
12 researchers found that DMSO can slowly release S2– ions into solution for synthesis
13
14
15 of S-based semiconductors under the facile one-pot solvothermal condition, resulting
16
17
18 in highly crystalline structures of Cu2S, CdS, ZnS and NiS.14-17 Compared to the other
19
20 S sources, DMSO is low-cost and easy to operate. We therefore are inspired to
21
22
23 explore a new route to synthesize S-doped TiO2 (S-TiO2) by using DMSO as the S
24
25
26 source.
27
28
29
30
31
32
33
34
35
36
37
38
39
40 Scheme 1. Schematic illustration of the formation process of S-TiO2 powders.
41
42
43
44
Specifically in this paper, we report a new, low-cost and facile method for the
45
46 synthesis of S-TiO2 through a solvothermal method, as shown in Scheme 1. Titanium
47
48
49 butoxide was used as the Ti source and DMSO served as both the S source and the
50
51
52
solvent during the reaction. The resulting powders were then calcined at 500 °C to
53
54 further crystallize the structure and to remove any surface attached organic species.
55
56
57 Energy-dispersive X-ray spectroscopy (EDX) and X-ray photoelectron spectroscopy
58
59
60
(XPS) solidly confirmed the presence of the sulfur element in the final product,

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 4 of 24

1
2
3
4 suggesting the generation of S-TiO2. More importantly, these S-TiO2 nanoparticles
5
6
7 exhibited more excellent and long-term stable photocatalytic performance for the
8
9
10 degradation of organic pollutants under indoor sunlight illumination than commercial
11
12 P25 TiO2. The catalyst still kept high photoactivity even after several months of
13
14
15 exposure to the indoor sunlight irradiation. This result suggests a new approach to
16
17
18 achieve stable and highly efficient solar light driven photocatalysts for water
19
20 purification.
21
22
23
24 EXPERIMENTAL SECTION
25
26
27 Materials. Dimethyl sulfoxide (DMSO, certified ACS, Fisher Scientific), Titanium
28
29 (IV) butoxide (97%, Sigma-Aldrich), All other chemicals were purchased from
30
31
32 Sigma-Aldrich without further purification before use. Milli-Q water was used
33
34
35 throughout our experiments.
36
37 Synthesis of the S doped TiO2 (S-TiO2) nanoparticles. To synthesize the S-doped
38
39
40 TiO2 powders, titanium butoxide (5.1 g, 0.015 mol) was added into the 50 mL DMSO
41
42
43 solvent under stirring. The solution was continuous stirred at room temperature for 30
44
45 min, then transferred into a 100 mL Teflon autoclave and held at 180 °C for 18 h.
46
47
48 After that, the precipitates were collected by centrifugation, washed with water and
49
50
51 ethanol, and then dried in an oven at 60 °C for 12 h. After that, the powder was
52
53 calcined at 500 °C for 4 h, resulting in a light yellow S-TiO2 sample.
54
55
56 Photocatalytic performance. For catalytic experiments, 20 mg of samples were
57
58
59 dispersed in a 10 mL rhodamine B (RhB, 22.5 mg L−1) or methylene blue (MB, 20 mg
60
L−1) solution, wherein a 20 mL cuvette was used as the reactor. A 150 W xenon arc
4

ACS Paragon Plus Environment


Page 5 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 lamp installed in a laboratory lamp housing system (LS 150 Xenon Arc Lamp Source,
5
6
7 Abet Technologies) was employed as the light source. The light passed through a
8
9
10 polyester Lee type-226 cut off filter (400nm) before entering the reactor. The reaction
11
12 system was kept for 30 min in dark to achieve an equilibrium adsorption state before
13
14
15 visible-light irradiation. The photodegradation of RhB and MB was investigated by
16
17
18 measuring the real-time UV−vis absorption of RhB and MB at 554 nm and 665 nm,
19
20 respectively. Aliquot of the reaction solution (0.25 mL) was taken out from the
21
22
23 reaction system for the real-time sampling. The pollutants relative concentrations
24
25
26 (C/C0) variation were used to evaluate the photocatalytic activities, where C was the
27
28 concentration of RhB or MB at a real-time t, and C0 was the concentration in the RhB
29
30
31 or MB solution before it was kept in dark. The integrated light intensity was measured
32
33
34 to be ca. 30 mW cm–2 by a visible–light radiometer (model: PM200, Thorlabs
35
36 GmbH).
37
38
39 In order to investigate the wavelength-dependent photocatalytic performance, the
40
41
42 incident light was passed the assigned bandpass filter (365 ±15 nm, 400±15 nm and
43
44 465 ±15 nm) before entering the reactor. In these cases, 50 mg of samples were
45
46
47 dispersed into 40 mL of RhB or MB aqueous solution (10 mg L-1) for the catalytic
48
49
50 experiments. The average intensity of the incident light was ca. 4.4 mW cm–2.
51
52 To investigate the long term photocatalytic activity, 20 mg of S-TiO2 samples
53
54
55 were dispersed in two cuvettes containing 10 mL RhB (13.5 mg L−1) and MB (15 mg
56
57
58 L−1) solution, respectively. The reaction system was kept beside the laboratory
59
60 window without stirring. The indoor sunlight (>350 nm) irradiated the reactor from

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 6 of 24

1
2
3
4 9:00 am to 3:00 pm during a sunny day. The light intensity of the indoor sunlight was
5
6
7 5 ~ 35 mW cm–2. When the dyes were degraded completely, the concentrated RhB
8
9
10 and MB solutions were added in the reactor to reach the initial concentration for a
11
12 new cycle. The cycling experiment was repeated for over three-month period to
13
14
15 evaluate the long term photocatalytic stability under indoor sunlight irradiation. The
16
17
18 recycled photocatalysts were then collected to further determine the photocatalytic
19
20 activity. In details, 20 mg of catalyst powders were dispersed in two cuvettes
21
22
23 containing 10 mL RhB (13.5 mg L−1) and MB (15 mg L−1) solution, respectively. The
24
25
26 solutions were irradiated under simulate indoor sunlight (λ>350 nm, 50 mW cm-2) for
27
28 24 minutes.
29
30
31 Apparatus and measurements. Transmission electron microscopy (TEM) studies
32
33
34 were conducted on a TECNAI-20 electron microscope operating at an accelerating
35
36 voltage of 200 kV. Scanning electron microscope (SEM, S–4700) was used to
37
38
39 determine the morphology of the as-prepared composite samples. The energy
40
41
42 dispersive X-ray (EDX) analysis was conducted with a Horiba EMAX X-act energy
43
44 dispersive spectroscope that was attached to the S-4700 system. The X–ray diffraction
45
46
47 (XRD) measurements were performed on a PANalytical X' Pert PRO MRD system
48
49
50 with Cu Ka radiation (k =1.54056 Å) operated at 40 kV and 30 mA. UV-vis diffuse
51
52 reflectance spectra were obtained on an UV-vis spectrophotometer (Hitachi, Model
53
54
55 U-3900) using BaSO4 as the reference. X-ray photoelectron spectroscopy (XPS) was
56
57
58 performed on an ESCALab220i-XL electron spectrometer from VG Scientific using
59
60 300 W Al Kα radiation. The binding energies were referenced to the C1s line at 284.8

ACS Paragon Plus Environment


Page 7 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 eV from adventitious carbon. All of these measurements were carried out at room
5
6
7 temperature. BET surface areas were measured on Autosorb-1 (Quantachrome Inc.)
8
9
10 using N2 adsorbent at 77K.
11
12
13 RESULTS AND DISCUSSION
14
15
16 The morphology of the as-synthesized S-TiO2 nanoparticles was analyzed by
17
18
19 scanning electron microscope (SEM) and transmission electron microscope (TEM).
20
21 As shown in Figure 1a and 1b the S-TiO2 nanoparticles appeared to be small
22
23
24 spherical particles with the average size ca 9.1 nm (See insert of the Figure 1b, the
25
26
27 corresponding size histogram of S-TiO2 nanoparticles counted from the TEM image).
28
29 The X-ray powder diffraction (XRD) pattern of the S-TiO2 powders was carried out to
30
31
32 investigate the crystalline phase of the as-prepared sample. Figure 2 clearly reveals
33
34
35 the peaks at 25.5°, 38.0°, 48.1°, 54.2°, 55.2°, 62.8°, 69.1°, 70.3°, and 75.3° which
36
37 were assigned to the diffraction of the (101), (004), (200), (105), (211), (204), (116),
38
39
40 (220), and (215) crystal planes, respectively, of anatase TiO2 (JCPDS No. 21-1272).13
41
42
43 As there are no diffraction peaks due to the rutile phase observed in the spectrum, we
44
45 conclude that the as-prepared S-TiO2 was in a purely anatase structure. Furthermore,
46
47
48 the average grain size determined from the Scherrer equation (D = 0.9λ/βcosθ) was
49
50
51 about 9.5 nm based on (101) diffraction peak,13 which also consisted with the TEM
52
53 image observation.
54
55
56
57
58
59
60

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 8 of 24

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Figure 1. (a) SEM and (b) TEM images of the as-prepared S-TiO2 nanoparticles. The
17
18
19 insert in image b is S-TiO2 size distribution histogram deduced from the TEM image.
20
21
22 The UV–Vis diffuse reflectance spectra are used to analyze the optical properties
23
24 and the bandgap energy of the samples. Figure 3A shows the results from the
25
26
27 reflectance measurements of S-TiO2 and commercial P25 TiO2 nanoparticles.
28
29
30 Compared to the P25 TiO2 (curve b, the absorption edge at ca. 398 nm), the
31
32 absorption edge of S-TiO2 red–shifted to ca. 435 nm (curve a), and the corresponding
33
34
35 UV-Vis spectrum shows a trailing absorption from 400 nm to 550 nm. The
36
37
38 photographic images (insets in Figure 3A) also show a distinct colour difference
39
40 between the as-prepared S-TiO2 nanoparticles (light yellow) and commercial P25 TiO2
41
42
43 nanoparticles (white). Extrapolation of the reflectance was used to obtain the band
44
45
46 gap energy of the samples. The bandgap energy of S-TiO2 and P25 TiO2 were ca. 2.85
47
48 eV and 3.1 eV, respectively, as shown in Figure 3B. The “tail-like” feature and
49
50
51 bandgap narrowing were attributed to the introduction of S atoms in the lattice of TiO2.
52
53
54 The formation of doping states can reduce the electron transition energy from the
55
56 valence to conduction band and thus lead to a red-shift of the absorption edge.
57
58
59
60

ACS Paragon Plus Environment


Page 9 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18 Figure 2. XRD pattern of the as-prepared S-TiO2 nanoparticles.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 Figure 3. (A) The UV–Vis diffuse reflectance spectra and (B) plots of (αhν)2 vs.
36
37
38 photon energy of S-TiO2 (a) and P25 TiO2 (b) nanoparticles. The inserts in image A
39
40 are the photographs of the S-TiO2 (a) and commercial P25 TiO2 (b) nanoparticles.
41
42
43 Table 1. Ti, O and S contents of as-prepared S-TiO2 were determined by EDX and
44
45 XPS study.
46
47 Element EDX (atom %) XPS (atom %)
48
49 Ti 32.07 31.17
50
51 O 66.24 66.83
52
53 S 1.69 2.00
54
55
56
57
58
59
60

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 10 of 24

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Figure 4. EDX elemental analysis of of the as-prepared S-TiO2 nanoparticles.
20
21
22 To verify the existence of S, the EDX and XPS spectra of S-TiO2 were applied
23
24 and shown in Figure 4 and Figure 5. First, from the EDX spectrum (Figure 4), Ti, O,
25
26
27 and S elements were observed in the as-prepared samples, suggesting the formation of
28
29
30 S-doped TiO2. Moreover, the XPS survey spectrum clearly shows the O 1s, Ti 2p, S
31
32 2p and C 1s core levels (Figure 5). Specifically, the O 1s XPS spectrum can be
33
34
35 resolved into two peaks at ca. 529.9 and 531.7 eV, which are ascribed to Ti–O and
36
37
38 surface OH species, respectively.18 The Ti 2p in S-TiO2 displays two peaks centered at
39
40 458.8 and 464.4 eV, which can be ascribed to the binding energy of Ti 2p3/2 and Ti
41
42
43 2p1/2, respectively.18 The presence of S was confirmed by a peak at 168.5 eV. This
44
45
46 peak can be further deconvoluted into two peaks at 168.5 eV and 169.7 eV, which can
47
48 be assigned to S 2p3/2 and S 2p1/2 respectively.11,13 Generally, the peak at ~168.5 eV is
49
50
51 assigned to the S6+ state, with a appearance of S 2p3/2 peak about twice the intensity or
52
53
54 area higher than that of S 2p1/2 peak.11 Hence, the S element might be S6+ in the lattice
55
56 of S-TiO2. This is also similar to previous literature reports.7,11,12 The C 1s XPS
57
58
59 spectrum showed one peak at 284.8 eV and a shoulder at around 288.8 eV, which are
60

assigned to C-C bonds and C-O bonds19, respectively. This is possibly due to the
10

ACS Paragon Plus Environment


Page 11 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 environmental species detected during the measurement. However, we did not
5
6
7 observe the formation of C-Ti bonds because of the missing of the bands at ~ 282 eV,
8
9
10 suggesting element C was not doped into the crystalline lattice20. The semiquantitative
11
12 analysis of the as-synthesized S-TiO2 by EDX and XPS are summarized in Table 1.
13
14
15 First, the results indicate that the atomic ratio between O and Ti is similar to the
16
17
18 theoretical stoichiometric atomic ratio, indicating the formation of TiO2. The S
19
20 content in S-TiO2 is about 2 atom% determined by EDX and XPS (Table 1). All the
21
22
23 above results prove the successful generation of an S doped TiO2 sample.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44 Figure 5. (a) XPS survey spectrum of S-TiO2 nanoparticles, (b) Ti 2p, (c) O 1s, (d) S
45
46
47 2p and (e) C 1s signals taken from S-TiO2.
48
49
50 The above analyses of the S-TiO2 suggest that the S element was successfully
51
52 introduced into the lattice of TiO2, resulting a distinct absorption in the visible light
53
54
55 range. This visible light response enables the S–TiO2 nanoparticles a potential to
56
57
58 utilize visible light in solar spectrum for catalytic degradation of organic pollutants.
59
60 We chose two typical dyes molecules of rhodamine B (RhB) and methylene blue (MB)

11

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 12 of 24

1
2
3
4 as targeting objects to evaluate the photocatalytic activities of the as-prepared S-TiO2.
5
6
7 Figure 6 shows the molecular structure of the above two dyes molecules.
8
9
10 Photo-degradations of RhB and MB under visible-light irradiation were performed as
11
12 the photoreaction probes to evaluate the photocatalytic activity of S-TiO2, and the
13
14
15 results are shown in Figure 7. For comparison, commercial P25 TiO2 was also used at
16
17
18 the same experimental conditions. There was negligible degradation of RhB and MB
19
20 pollutants when no catalysts were used after 120 min of visible-light irradiation.
21
22
23 Around 82.8% and 63.8% of RhB and MB molecules degraded in 120 min when P25
24
25
26 TiO2 was used. In contrast, when our S-TiO2 nanoparticles were used as the
27
28 photocatalyst, the photoactivity was significantly improved, and the RhB and MB
29
30
31 molecules were degraded nearly 97% and 100% under visible-light irradiation in 120
32
33
34 minutes. Moreover, the BET surface area of the as-prepared S-TiO2 is ca. 124.24 m2
35
36 g-1, which is larger than the P25 TiO2 (generally is 35~65 m2 g-1, from Sigma-Aldrich).
37
38
39 The larger surface area might also play a role to improve the catalytic activity.
40
41
42 Furthermore, we also synthesized S-doped TiO2 following the previously
43
44 reported work by using thiourea as S source through the precipitation method (details
45
46
47 see Supporting Information)7, and the obtained samples was named as S-TiO2-P.
48
49
50 However, there were only 62.3% and 72.4% of RhB and MB molecules degraded
51
52 respectively in 120 min over S-TiO2-P sample at the same experiment conditions.
53
54
55 (Figure S1). The results clearly suggest by using DMSO as both S source and solvent
56
57
58 is a better method to fabricate high efficient visible-light active S-TiO2 photocatalyst
59
60 over the other reported work.

12

ACS Paragon Plus Environment


Page 13 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4
5
6
7
8
9
10
11
12
13
14 Figure 6. Molecular structures of the RhB and MB molecules.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Figure 7. (a and d) Photocatalytic activities, and (b and e) kinetic linear simulation
37
38
39 curves for the degradation of RhB (a and b) and MB (d and e) pollutants without
40
41
42 catalysts and with commercial P25 TiO2 and S-TiO2 catalysts under visible light
43
44 (λ>400 nm) irradiation. The real-time absorption spectra of RhB (c) and MB (f)
45
46
47 solution during the photodegradation process over S-TiO2 under visible-light
48
49
50 illumination from 0 min to 120 min.
51
52 As plotted in Figure 7b and 7e, there is a nice linear correlation between ln(C/C0)
53
54
55 and the reaction time (t), indicating that the decomposition of RhB and MB over TiO2
56
57
58 photocatalysts follows the first-order kinetics. The rate constants of the catalytic
59
60 degradation of RhB and MB over commercial TiO2 were 0.014 min-1 and 0.008 min-1,

13

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 14 of 24

1
2
3
4 respectively, while those of S-TiO2 were 0.028 min-1 and 0.043 min-1. The
5
6
7 photocatalytic efficiency of S-TiO2 was improved ca. 2 and 5 times compared with
8
9
10 the P25 TiO2 for the degradation of RhB and MB molecules, respectively. The reason
11
12 of the improved catalytic performance of the as-prepared S-TiO2 is attributed to the S
13
14
15 element introduced into the lattice of TiO2, which narrows the bandgap of the TiO2,
16
17
18 resulting in enhanced visible light absorption of solar energy.
19
20 The solvothermal temperature controls the DMSO decomposition, and further
21
22
23 controls the S-doping ratio in the TiO2 crystalline lattice. As DMSO is non-degradable
24
25
26 below 150 oC, and decomposes at boiling point 189 oC, possibly leading to the
27
28 explosion,17 the solvothermal reaction was controlled between 140 oC and 180 oC. We
29
30
31 did not observe S-doping at 140 0C possibly due to the DMSO non-degradation.
32
33
34 Figure S2 shows the amount of S doping in TiO2 was ca. 0.73 when the solvothermal
35
36 treatment temperature was 160 °C (S-TiO2-160). Figure S3 shows RhB and MB
37
38
39 molecules were degraded nearly 71.6% and 87.3% under visible-light irradiation
40
41
42 within 120 min by using S-TiO2-160 photocatalysts. S-TiO2-160 shows the poor
43
44 photocatalytic activities compared with S-TiO2 synthesized at 180 °C solvothermal
45
46
47 treatment, possibly due to the lower S-doping in the lattice.
48
49
50 Calcination temperature determines the crystallinity of the formed TiO2, and the
51
52 Figure S4 shows the XRD patterns of the as-prepared S-TiO2 calcined at 400 °C,
53
54
55 500 °C and 600 °C. The S-TiO2 presented a pure anatase structure at 400 °C and
56
57
58 500oC calcination, and a mixture of anatase and rutile phases at 600 oC. Figure 8
59
60 shows the photocatalytic degradation of RhB and MB pollutants by using S-TiO2

14

ACS Paragon Plus Environment


Page 15 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 calcined at different temperature. The catalytic performance for S-TiO2 calcined at
5
6
7 400 oC and 500 oC was similar, but decreased when the calcination temperature was
8
9
10 increased to 600 °C. Apparently, the catalytic performance is correlated with the
11
12 weighting percent of the active photocatalytic component anatase structure. High
13
14
15 temperature (600 oC) calcination promotes the transition from anatase to rutile
16
17
18 structure, deteriorating the catalytic performance accordingly.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 8. Photocatalytic activities of various S-TiO2 by different calcinations
39
40 temperature for the degradation of RhB and MB under visible light irradiation with
41
42
43 120 min.
44
45
46 Different bandpass filters were used to study the irradiation light wavelength
47
48 dependence during the photocatalytic degradation of RhB and MB pollutants, and the
49
50
51 results were shown in Figure 9. Commercial available P25 TiO2 was used as a
52
53
54 reference for comparison. In UV band when 365±15 nm bandpass filter was used, the
55
56 as-prepared S-TiO2 and P25 TiO2 showed the similar photocatalytic activities. In
57
58
59 visible band 400±15 nm, the photocatalytic activity of S-TiO2 was more than two
60

times higher than that of P25 TiO2. There was very weak dyes degradation on both
15

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 16 of 24

1
2
3
4 S-TiO2 and P25 TiO2 when the longer wavelength bandpass filter 465±15 nm was
5
6
7 used. These results first are consistent with the bandgap analyzation as discussed in
8
9
10 Figure 3; furthermore, by comparing the photocatalytic activity of P25 TiO2 and
11
12 S-TiO2, the activity similarity at 365±15 nm band and apparent increase at 400±15 nm
13
14
15 for S-TiO2 suggest that the main reason for enhancing catalytic activities were
16
17
18 attributed to S element introduced into the lattice of TiO2 other than the increased
19
20 surface area. In summary, S-TiO2 has superior photocatalytic performance than P25
21
22
23 TiO2 to degrade the organic pollutants.
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47 Figure 9. Wavelength dependence for photocatalytic degradation of RhB and MB by
48
49
50 using commercial P25 TiO2 and S-TiO2 as photocatalysts. The light exposure time
51
52 was 4 hours.
53
54
55 Besides catalytic performance, stability is another important factor for high
56
57
58 quality catalysts with practical applications. Moreover, in a real environment, sunlight
59
60 is the most economical and renewable energy source to explore for the application of

16

ACS Paragon Plus Environment


Page 17 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 photocatalysts. Considering the above issues, the stability of S–TiO2 photocatalyst
5
6
7 was investigated by recycling degradation of RhB and MB pollutants under indoor
8
9
10 sunlight (λ>350 nm) irradiation. At the end of each cycle, the organic pollutants were
11
12 resupplied to the initial concentration for the next run. It should be noted that these
13
14
15 experiments were carried out without stirring or extra light irradiation. Figure 10
16
17
18 shows the catalyst still kept a high catalytic capability to degrade both RhB and MB
19
20 over more than 40 cycles in a 3 month test span by using S-TiO2 as the photocatalyst.
21
22
23 We estimated the number of the bonded surface hydroxyl group on S-TiO2 was the
24
25
26 same as that of P25 TiO2, which is about 9.5 1019/g (equals 1.58 104 mol/g) as
27
28 previously reported, 21 and all of surface bonded –OH groups acted as active sites for
29
30
31 catalytic reaction. As a result, if we regard the target organic molecule as a catalysant,
32
33
34 the turnover number for RhB and MB catalytic degradation is 3.66 and 6.08,
35
36 respectively.
37
38
39 To further compare the photocatalytic performance between S-TiO2 and P25
40
41
42 TiO2, we investigated the catalytic activity for the degradation of RhB and MB
43
44 pollutants before and after the long term irradiation. All the photodegradation
45
46
47 reactions were carried out under simulated indoor sunlight (λ>350 nm) irradiation for
48
49
50 24 min. As shown in Figure S5, the fresh catalysts S-TiO2 and P25 TiO2 used before
51
52 the long-term irradiation treatment both exhibited a high photoactivity, and the RhB
53
54
55 and MB were degraded nearly 98.5% and 100%, 94.6 % and 94.4 %, respectively.
56
57
58 P25 TiO2 showed significant activity drop after cycling 20 times under simulate
59
60 indoor sunlight irradiation, and the photocatalytic degradation of RhB and MB was

17

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 18 of 24

1
2
3
4 reduced to 77.2 % and 75 %. On the other hand, the S-TiO2 still displayed high
5
6
7 catalytic activity even after long-term indoor sunlight irradiation. Figure S5A shows
8
9
10 that 96.2 % of RhB and and 93.1 % MB were degraded in 24 minutes after S-TiO2
11
12 were recycled 40 times. These results solidly suggest that as-synthesized S-TiO2 could
13
14
15 be employed as a long term stable and efficient catalyst for the water purification
16
17
18 under solar light illumination in practical use.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39 Figure 10. Long-term (3 months) continuous photocatalytic degradation of RhB and
40
41
42 MB pollutants by using S-TiO2 catalysts (20 mg) under indoor sunlight (λ>350 nm)
43
44 irradiation.
45
46
47 The high photocatalytic performance of S-TiO2 is first owing to S doping in the
48
49
50 Ti-O-Ti crystalline structure, resulting in a red-shift of the absorption edge and
51
52 narrowed bandgap, and leading to the visible light photosensitivity. Furthermore, XPS
53
54
55 results (Figure 5) suggest sulphur is in +6 state, hence when S atom takes the place of
56
57
58 Ti, it forms SO4 tetrahedral unit in the lattice. Visible light illumination on the
59
60 photocatalyst forms the photogenerated pairs. Compared with photogenerated holes

18

ACS Paragon Plus Environment


Page 19 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 (TiO2+), (SO2+) holes are expected to have higher oxidation potential and can easily
5
6
7 decompose any unreacted intermediate species adsorbed on the surface. This strong
8
9
10 surface self-cleaning property leads to long-term photocatlytic stability.
11
12 CONCLUSION
13
14
15 In conclusion, visible-light-sensitive S-doped TiO2 with high photocatalytic
16
17
18 performance and long-term stability was synthesized through a facile solvothermal
19
20 method, in which DMSO acted as both the solvent and the S source. The existence of
21
22
23 S in the lattice of TiO2 resulted in a narrowing of the S-TiO2 bandgap and therefore
24
25
26 providing a visible-light catalytic response. Compared with commercial P25 TiO2,
27
28 S–TiO2 showed evidently enhanced photoactivity for the degradation of RhB and MB
29
30
31 under visible light irradiation. Excitingly, S-TiO2 displayed excellent catalytic
32
33
34 stability after long-term indoor sunlight irradiation. This work provides a new method
35
36 for developing stable and efficient visible-light-driven photocatalysts to degrade
37
38
39 organic pollutants. These materials would also be expected to have promising
40
41
42 applications in solar cells, water splitting, and other light harvesting systems.
43
44
45 ASSOCIATED CONTENT
46
47
48 Supporting Information
49
50
51 The Supporting Information is available free of charge on the ACS Publications
52
53 website
54
55
56 Preparation of S-doped TiO2 by using thiourea as S source (S-TiO2-P) and
57
58
59 corresponding photocatalytic performance, EDX spectrum and photocatalytic
60

19

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 20 of 24

1
2
3
4 activities of S-TiO2-160, XRD patterns of S-TiO2 annealed at different temperature,
5
6
7 and photocatalytic stability of S-TiO2 and P25 TiO2
8
9
10
11 AUTHOR INFORMATION
12
13 Corresponding Author
14
15
16 * Tel./Fax: +1-416-9784526. E-mail: clu@chem.utoronto.ca (C. Lu);
17
18
19 cgoh@chem.utoronto.ca (M. C. Goh).
20
21 Author Contributions
22
23
24 The manuscript was written through contributions of all authors. All authors
25
26
27 have given approval to the final version of the manuscript.
28
29 Notes: The authors declare no competing financial interest.
30
31
32
33 ACKNOWLEDGMENT
34
35
36 Funding for this work was provided by the Natural Sciences and Engineering
37
38 Research Council of Canada.
39
40
41
42 REFERENCES
43
44
45 (1) Chen, X.; Mao, S. S. Titanium dioxide nanomaterials:  synthesis, properties,
46
47
modifications, and applications. Chem. Rev. 2007, 107, 2891-2959.
48
49
50 (2) Kapilashrami, M.; Zhang, Y.; Liu, Y.-S.; Hagfeldt, A.; Guo J. Probing the optical
51
52
53 property and electronic structure of TiO2 nanomaterials for renewable energy
54
55
applications. Chem. Rev. 2014, 114, 9662 -9707.
56
57
58 (3) Fujishima, A.; Zhang, X.; Tryk, D. A.; TiO2 photocatalysis and related surface
59
60
phenomena. Surf. Sci. Rep. 2008, 63, 515-582.

20

ACS Paragon Plus Environment


Page 21 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 (4) Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-doped titanium dioxide as
5
6
7 visible-light-sensitive photocatalyst: designs, developments, and prospects. Chem.
8
9
10 Rev. 2014, 114, 9824-9852.
11
12 (5) Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light
13
14
15 photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269-271.
16
17
18 (6) Umebayashi, T.; Yamaki, T.; Itoh, H.; Asai, K. Band gap narrowing of titanium
19
20 dioxide by sulfur doping. Appl. Phys. Lett. 2002, 81, 454-456.
21
22
23 (7) Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; Matsumura, M.
24
25
26 Preparation of S-doped TiO2 photocatalysts and their photocatalytic activities
27
28 under visible light. Appl. Catal. A: Gen. 2004, 265, 115-121.
29
30
31 (8) Ho, W.; Yu, J. C.; Lee, S. Low-temperature hydrothermal synthesis of S-doped
32
33
34 TiO2 with visible light photocatalytic activity. J. Solid State Chem. 2006, 179,
35
36 1171-1176.
37
38
39 (9) Li, H.; Zhang, X.; Huo, Y.; Zhu, J. Supercritical preparation of a highly active
40
41
42 S-doped TiO2 photocatalyst for methylene blue mineralization. Environ. Sci.
43
44 Technol. 2007, 41, 4410-4414.
45
46
47 (10) Chen, X.; Burda, C. The electronic origin of the visible-light absorption
48
49
50 properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 2008,
51
52 130, 5018-5019.
53
54
55 (11) Periyat, P.; Pillai, S. C.; McCormack, D. E.; Colreavy, J.; Hinder, S. J. Improved
56
57
58 high-temperature stability and sun-light-driven photocatalytic activity of
59
60 sulfur-doped anatase TiO2. J. Phys. Chem. C 2008, 112, 7644-7652.

21

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 22 of 24

1
2
3
4 (12) Tian, G.; Pan, K.; Fu, H.; Jing, L.; Zhou, W. Enhanced photocatalytic activity of
5
6
7 S-doped TiO2–ZrO2 nanoparticles under visible-light irradiation. J. Hazardous
8
9
10 Mater. 2009, 166, 939-944.
11
12 (13) Naik, B.; Parida, K. M.; Gopinath, C. S. Facile synthesis of N- and
13
14
15 S-incorporated nanocrystalline TiO2 and direct solar-light-driven photocatalytic
16
17
18 activity. J. Phys. Chem. C 2010, 114, 19473-19482.
19
20 (14) Wu, Z.; Pan, C.; Yao, Z.; Zhao, Q.; Xie, Y. Large-scale synthesis of single-crystal
21
22
23 double-fold snowflake Cu2S dendrites. Cryst. Growth Des. 2006, 6, 1717-1719.
24
25
26 (15) Cao, A.; Liu, Z.; Chu, S.; Wu, M.; Ye, Z.; Cai, Z.; Chang, Y.; Wang, S.; Gong, Q.;
27
28 Liu, Y. A facile one-step method to produce graphene–CdS quantum dot
29
30
31 nanocomposites as promising optoelectronic materials. Adv. Mater. 2010, 22,
32
33
34 103-106.
35
36 (16) Li, Y.; Liu, Y.; Shen, W.; Yang, Y.; Wen, Y.; Wang, M. Graphene–ZnS quantum
37
38
39 dot nanocomposites produced by solvothermal route. Mater. Lett. 2011, 65,
40
41
42 2518-2521.
43
44 (17) Wang, W.; Wang, S.-Y.; Gao, Y.-L.; Wang, K.-Y.; Liu, M. Nickel sulfide
45
46
47 nanotubes formed by a directional infiltration self-assembly route in AAO
48
49
50 templates. Mater. Sci. Engineering: B 2006, 133, 167-171.
51
52 (18) Zhai, C.; Zhu, M.; Lu, Y.; Ren, F.; Wang, C.; Du, Y.; Yang, P. Reduced graphene
53
54
55 oxide modified highly ordered TiO2 nanotube arrays photoelectrode with
56
57
58 enhanced photoelectrocatalytic performance under visible-light irradiation. Phys.
59
60 Chem. Chem. Phys. 2014, 16, 14800-14807.

22

ACS Paragon Plus Environment


Page 23 of 24 ACS Sustainable Chemistry & Engineering

1
2
3
4 (19) Yang, X.; Cao, C.; Hohn, K.; Erickson, L.; Maghirang, R.; Hamal, D.; Klabunde,
5
6
7 K. Highly visible-light active C- and V-doped TiO2 for degradation of
8
9
10 acetaldehyde. J. Catal. 2007, 252, 296–302.
11
12 (20) Magnuson, M.; Lewin, E.; Hultman, L.; Jansson, U., Electronic structure and
13
14
15 chemical bonding of nanocrystalline-TiC/amorphous-C nanocomposites. Phys.
16
17
18 Rev. B 2009, 80, 235108.
19
20 (21) Takahashi, J.; Itoh, H.; Motai, S.; Shimada, S., Dye adsorption behavior of
21
22
23 anatase- and rutile-type TiO2 nanoparticles modified by various heat-treatments.
24
25
26 J. Mater. Sci. 2003, 38, 1695-1702.
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

23

ACS Paragon Plus Environment


ACS Sustainable Chemistry & Engineering Page 24 of 24

1
2
3
4 For Table of Contents Use Only
5
6
7
Title: A New Method to Synthesize S-Doped TiO2 with Stable and Highly Efficient
8
9
Photocatalytic Performance under Indoor Sunlight Irradiation
10
11 Authors: Mingshan Zhu, Chunyang Zhai, Liqun Qiu, Cheng Lu, Andrew Paton,
12
13 Yukou Du, and M. Cynthia Goh
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35 A new facile method for the synthesis of S-doped TiO2 with excellent and long-term
36
37 stable photocatalytic performance under indoor sunlight irradiation
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

24

ACS Paragon Plus Environment

Вам также может понравиться