Nanofibers Modified by Graphitic Carbon Nitride ... - ACS Publications

Jun 2, 2016 - Institute of Engineering, Tribhuvan University, Kathmandu, Nepal. •S Supporting Information. ABSTRACT: We report a direct approach to ...
0 downloads 0 Views 6MB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article

Electrospinning directly synthesized porous TiO2 nanofibers modified by graphitic carbon nitride sheets for enhanced photocatalytic degradation activity under solar light irradiation Surya Prasad Adhikari, Ganesh Prasad Awasthi, Han Joo Kim, Chan-Hee Park, and Cheol Sang Kim Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b01085 • Publication Date (Web): 02 Jun 2016 Downloaded from http://pubs.acs.org on June 3, 2016

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.

Langmuir 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 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 1: Schematic illustration for the synthesis of the TiO2/g-C3N4 composite. 48x16mm (600 x 600 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 2: XRD patterns of (a) g-C3N4, (b) TiO2, (c) TCN-1, (d) TCN-3, (e) TCN-5 and (f) TCN-10. Inset XRD pattern showing the 23-30 degree region. 116x96mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 2 of 70

Page 3 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 3: Raman shift of TiO2 and TCN-5. 99x70mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 4: FT-IR images of (a) g-C3N4, (b) TiO2, (c) TCN-1, (d) TCN-3, (e) TCN-5 and (f) TCN-10. Inset FTIR spectra showing 2500-4000 cm-1 region. 126x113mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 4 of 70

Page 5 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 5: FE-SEM images of (a) TTIP/PVAC fibers before calcinations, (b) TiO2 fibers after calcination, (c) TTIP/PVAC/g-C3N4 fibers before calcinations, and (d) TCN composite after calcinations. 95x65mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 6: (a) TEM images of (a) TiO2 NFs, (b) TCN, and (c) HR-TEM images of TCN composite. 118x100mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 6 of 70

Page 7 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 7: (a) XPS survey spectrum and high resolution XPS spectra of (b) C1s, (c) N1s, (d) Ti 2p, (e) O1s and (f) shifting of Ti2p peak in TCN-5 composite. 160x182mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 8: (a) UV-vis absorption spectra of TiO2, g-C3N4 and TCN-x samples and (b) plots of (αhν)2 vs. photon energy (hν) for the band gap energies of different samples. 53x20mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 8 of 70

Page 9 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 9: (a) Photodegradation of RhB under natural sun light in the corresponding time intervals; (b) Photo images showing the initial 10-ppm RhB solution and the remaining RhB in the solution after being irradiated for 40 min; (c) Photodegradation of RB-5 under natural sunlight in the corresponding time intervals; and (d) Photo images showing the initial 10- ppm RB-5 solution and the remaining RB-5 in the solution after being irradiated for 20 min. 86x53mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 10: Kinetics of photocatalytic degradation of different catalysts, (a) 10-ppm RhB solution, and (b) 10ppm RB-5 solution. 57x23mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 10 of 70

Page 11 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 11: PL spectra of g-C3N4, TiO2 and TCN-5 111x89mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 12: Cyclic voltammograms of TiO2, g-C3N4 and TCN-5 102x75mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 12 of 70

Page 13 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 13: Electrochemical impedance spectra of TiO2, g-C3N4 and TCN-5 107x81mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure 14: Schematic diagram of photogenerated charges in the TCN hetero-junction. 120x104mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 14 of 70

Page 15 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure 15: Cyclic runs of TCN-5 in the photocatalytic degradation over RhB under natural light irradiation. 74x39mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure S1: FE-SEM images of g-C3N4 before and after sonication (a) and (b), respectively, SEM-EDX image with real atomic compositions (c) and (d), and XPS survey spectrum with constituent elements of g-C3N4. 110x87mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 16 of 70

Page 17 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure S2: Porous formation mechanism of TiO2 NFs 44x14mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure S3: FE-SEM images of electrospun PVAc NFs with g-C3N4 sheets in different magnifications. 108x83mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 18 of 70

Page 19 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure S4: High resolution XPS spectra of Ti 2p, O1s, C1s, and N1s for the pristine g-C3N4 and TiO2 NFs. 105x79mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure S5: Binding energy shifting comparison of O1s high resolution XPS spectra from pristine TiO2 NFs and TCN-5 composite. 204x299mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 20 of 70

Page 21 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Figure S6: Distribution of pore diameter 103x76mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Figure S7: Life time comparison of g-C3N4 sheets, TiO2 NFs and TCN-5 from FT-IR spectra before and after 4 cyclic runs 87x55mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 22 of 70

Page 23 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

Table S1: BET surface area and Langmuir surface area of different samples 42x12mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Table S2: Porous characteristics of g-C3N4, TiO2 and TCN-5 samples 76x41mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 24 of 70

Page 25 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

Electrospinning directly synthesized porous TiO2 nanofibers modified by graphitic carbon

2

nitride sheets for enhanced photocatalytic degradation activity under solar light irradiation

3

Surya Prasad Adhikari1, 2, 3, Ganesh Prasad Awasthi1, Han Joo Kim4, Chan Hee Park1, 3*, Cheol Sang

4

Kim1, 3*

5

1

6

Republic of Korea

7

2

Institute of Engineering, Tribhuvan University, Kathmandu, Nepal

8

3

Division of Mechanical Design Engineering, Chonbuk National University, Jeonju 561-756,

9

Republic of Korea

Department of Bionanosystem Engineering, Chonbuk National University, Jeonju 561-756,

10

4

11

University, Jeonju 561-756, Republic of Korea

12

*Corresponding authors:

13

Tel.: +82-63-270-4284; Fax: +82-63-270-2460

14

E-mail:[email protected] (C. H. Park),

15

[email protected] (C. S. Kim)

Division of Convergence Technology Engineering, Engineering College, Chonbuk National

16 17 18 19 20 21 22 23 24

1

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 26 of 70

1

Abstract

2

We report a direct approach to the fabrication of a composite made of porous TiO2 nanofibers (NFs)

3

and graphitic carbon nitride (g-C3N4) sheets, by means of an angled two-nozzle electrospinning

4

combined with calcination process. Different wt % amounts of g-C3N4 particles in a polymer

5

solution from one nozzle, and TiO2 precursors containing the same polymer solution from another

6

nozzle, were electrospun and deposited on the collector. Structural characterizations confirm a well-

7

defined morphology of the TiO2/g-C3N4 composite in which the TiO2 NFs are uniformly attached

8

on the g-C3N4 sheet. This proper attachment of TiO2 NFs on the g-C3N4 sheets occurred during

9

calcination. The prepared composites showed the enhanced photocatalytic activity over the

10

photodegradation of rhodamine B and reactive black 5 under natural sunlight. Here, the synergistic

11

effect between the g-C3N4 sheets and the TiO2 NFs having anisotropic properties enhanced the

12

photogenerated electron-hole pair separation and migration, which was confirmed by the

13

measurement of photoluminescence spectra, cyclic voltammograms, and electrochemical

14

impedance spectra. The direct synthesis approach that is established here for such kinds of sheet-

15

like structure and porous NFs composites could provide new insights for the design of high-

16

performance energy conversion catalysts.

17 18

Keywords: Electrospinning, g-C3N4, Porous TiO2 NFs, Photocatalysis

19 20 21 22 23 24

2

ACS Paragon Plus Environment

Page 27 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

Introduction

2

The semiconductors photocatalysis for degrading organic pollutants is one of the most potentially

3

important technologies that have been given serious consideration in the past few decades.

4

Therefore, a large number of semiconductor materials like metal oxides and sulfides such as TiO2,

5

ZnO, ZnS, CdS and WO3 were successfully prepared by a variety of methods and many seek to

6

further maximize their efficiency

7

most promising and efficient energy conversion catalyst for controlling organic pollutants because

8

of its unique characteristics, including low cost, encouraging optical and electronic properties,

9

nontoxicity, and high chemical and photochemical stability

1, 2, 3, 4, 5

. Among these, TiO2 has been considered as one of the

6, 7, 8

. Moreover, among various TiO2

10

nanostructures, the porous one-dimensional (1D) morphologies of the NFs have drawn a significant

11

amount of attention due to their large surface area and distinctively-oriented NFs array, which

12

improves the separation efficiency for the charge carrier by reducing the recombination rate. In

13

recent years, many researchers have studied the relationship between porous TiO 2 NFs and

14

photocatalytic performance, and have reported that TiO2 NFs having porous structures enhanced the

15

photocatalytic performance. H. Hou et al. prepared porous TiO2 NFs from electrospinning and

16

demonstrated that the (1D) porous fibers exhibit the highest photocatalytic activity compared to

17

nanostructures composed of nanowires and naoparticles 9. Similarly, S. K. Choi et al. prepared

18

electrospun, mesoporous TiO2 NFs and showed that the enhanced photocatalytic activity was due to

19

the effect of porosity and NFs alignment, which could reduce the recombination of electron-hole (e-

20

h) pairs through interparticle charge transfer along the NFs framework

21

properties, practical application of TiO2 is severely limited because of the narrow range of effective

22

wavelengths which belongs to the UV range and also the too-fast recombination of photo-induced

23

electron-hole (e-h) pairs. Considering these limitations, a lot of effort has been devoted to

24

synthesizing

hetero-structured

composites

of

TiO2

with

3

ACS Paragon Plus Environment

other

10

. In spite of these

visible-light-responsive

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 28 of 70

1

semiconductors by this means extending the absorption edge of TiO2 to the visible range, and

2

reducing the recombination of photoinduced e-h pairs 11, 12, 13, 14.

3

Among various visible-light-responsive semiconductors, polymeric g-C3N4 has been utilized as a

4

coupled semiconductor in the most recent times. Because of its narrow band gap and outstanding

5

mechanical, thermal, and optical properties, this metal-free photocatalyst is showing massive

6

potential for application to the photocatalytic degradation under visible light. However, in spite of

7

all these properties, the photocatalytic activity of pure g-C3N4 is also limited due to the minimal

8

surface area, low visible-light utilization rate and high recombination rate of photogenerated e-h

9

pairs

15

. Therefore, many attempts were made to improve the photocatalytic properties of g-C3N4,

10

such as the introduction of a mesoporous structure 16, 17, element doping with metals or nonmetals 18,

11

19, 20, 21

12

combining π structures of g-C3N4 with highly active semiconductors has attracted much interest.

13

Therefore, blending TiO2 with g-C3N4 sheets could be an appropriate technique for reducing the

14

recombination of e-h pairs beyond those of TiO2 and g-C3N4. To date, several authors have invoked

15

different techniques for synthesizing TiO2/g-C3N4 composites. Recently, Y. Li. et al. reported a

16

TiO2/g-C3N4 hybrid from a seed-induced solvothermal method that enabled the fabrication of TiO2

17

nanostructures on g-C3N4 surfaces with enhanced photocatalytic activity under visible light

18

et al. prepared a g-C3N4-TiO2 hybrid Z-scheme photocatalyst by a facile calcination route

19

space separation efficiency of the photogenerated e-h pairs was enhanced through the Z-scheme

20

route. Mostly, g-C3N4-based hybrid photocatalysts, including g-C3N4/TiO2, were prepared

21

calcinations of precursor mixtures that yielded composite NPs that are prone to aggregate

22

practical applications of g-C3N4 and TiO2 inexorably depend on the morphology of the material.

23

Therefore, it is still a big challenge to control the structures of these photocatalysts. Previous

24

literature has reported that the sheet-like structures of g-C3N4-based composites exhibit distinctive

, and building hetero-structures

22, 23

. Among them, constructing hetero-structures by

4

ACS Paragon Plus Environment

24

25

. Yu.

. The

26

. The

Page 29 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

optical and electronics properties compared to bulk g-C3N4 composites. P. Niu et al. revealed that

2

the sheet-like structure of g-C3N4 exhibits increased lifetimes of photoexcited charge carriers

3

This improved photoexcited charge carrier property was due to the quantum confinement effect.

4

Similarly, among various TiO2 nanostructures, the one-dimensional TiO2 NFs or nanorods attracted

5

serious interest in view of the dimensional confinement and minimal agglomeration compared to

6

nanoparticles (NPs)

7

composite made of sheet-like g-C3N4 and porous TiO2 NFs will exhibit encouraging properties,

8

tending toward-enhanced the photocatalytic activity. So, we were inspired to develop a simple

9

technique for creating such a composite material.

28

27

.

. Therefore, bearing in mind these observations, it is to be expected that a

10

To our best knowledge, there has been no report on the fabrication of porous TiO2 NFs grafted onto

11

g-C3N4 sheets. Recently, Han C. et al. synthesized a composite of TiO2 NFs and g-C3N4 sheets by

12

electrospinning, in which g-C3N4 was embedded inside the NFs

13

time, TiO2 porous NFs-having sufficient length were dispersed homogeneously and attached on the

14

surfaces of g-C3N4 sheet, with minimal agglomeration of NFs from an angled two-nozzle

15

electrospinning process with calcinations. Here, the enormous surface area of g-C3N4 sheets

16

provides adequate surfaces for the direct attachment of TiO2 NFs, which make charge separation

17

more efficient. Such NFs have a better chance than NPs to be uniformly attached onto g-C3N4

18

sheets to form bonding, simply based on geometric considerations. Here also, a series of TiO2/g-

19

C3N4 with different g-C3N4 loadings with respect to the polymer weight have been prepared. The

20

photodegradation behaviors of rhodamine B (RhB) and reactive black (RB-5) over the prepared

21

photocatalyst were measured under sun light. The composite exhibits higher photoactivity than pure

22

TiO2 and g-C3N4 under natural sun light irradiation.

23

Thus, this proposed method provides a facile and straightforward approach for affixing porous TiO2

24

NFs on the surface of g-C3N4 sheets from a simple and low cost electrospinning process. 5

ACS Paragon Plus Environment

29

. But, in our study, for the first

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 30 of 70

1

Experimental

2

Materials

3

Commercially available g-C3N4 particles nicanite®, (Carbedon, Finland), acetic acid, titanium

4

isopropoxide (TTIP), polyvinyl acetate (PVAc, Mw = 500000) and N,N-Dimethylformamide (DMF)

5

were purchased from Sigma Aldrich and used as-received.

6

Fabrication of TiO2 NFs-intercalated g-C3N4 sheets

7

Here, TiO2/g-C3N4 hybrid photocatalyst was directly prepared by using an angled two-nozzle

8

electrospinning-calcination process. Typically, two different solutions, one containing g-C3N4 and

9

the other containing TiO2 precursors were made from the same polymer solution. First, PVAc

10

solution (18 wt %) in DMF was prepared by overnight magnetic stirring at room temperature.

11

Thereafter, g-C3N4 particles (1, 3, 5 and 10 wt % with respect to the weight of the PVAc) were

12

added to the PVAc solution and the mixture was subjected to bath sonication for two hours, to

13

disperse the g-C3N4 particles. Similarly, the TiO2 precursor-containing PVAc solution was prepared

14

by mixing 6 g of the PVAc solution and 5 g of clear solution of TTIP (obtained by dropwise

15

addition of acetic acid with continuous stirring). For two-nozzle electrospinning, one syringe

16

contained PVAc solution with g-C3N4 particles while the other contained PVAc solution with TiO2

17

precursors. The angle between the two nozzle tips was maintained at 80°. A schematic diagram for

18

the preparation of the composite is as shown in Figure 1. Electrospinning was carried out in room

19

conditions where the parameters include 18 kV of applied voltage, a tip-to-collector distance of 12

20

cm, and solution feed rate of 1 ml/h. The obtained electrospun mats were vacuum dried at 80 °C

21

overnight. Then the vacuum-dried electrospun mats were treated in air at 500 °C for 3h at the

22

heating rate of 5 °C/min. At such a high temperature, TTIP [Ti{OCH(CH] molecules could

23

decompose into TiO2, CO2 and H2O. The PVAc component was eliminated in the mean time. Here,

24

CO2 and H2O escaped rapidly, and only TiO2 and g-C3N4 sheets remained. At high temperature, 6

ACS Paragon Plus Environment

Page 31 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

TiO2 molecules could react and generate anatase/rutile-TiO2. Here, the composite material in which

2

TiO2 NFs attached on the surface of the g-C3N4 sheets are denoted as “TCN-x”, where, “TCN”

3

refers TiO2/g-C3N4 composite and x refers the wt % of g-C3N4 with respect to the polymer weight,

4

namely 1, 3, 5 and 10. Note that the mass ratio of g-C3N4 sheet affects the spinnability of the

5

solution. Above 10 wt % of g-C3N4 sheet, spinnability diminishes and it is difficult to obtain fibers

6

due to the increasing viscosity. Moreover, pristine TiO2 NFs were prepared the same as TiO2/g-

7

C3N4 from the solution containing TiO2 precursors, using a single-nozzle electrospinning process

8

with all conditions identical to that of the two-nozzle electrospinning.

9

Characterization

10

The X-Ray diffraction patterns of the resulting samples were obtained by using a Rigaku X-ray

11

diffractometer (XRD, Rigaku, Japan). Raman shift was investigated by Raman Spectroscopy

12

(Nanofinder® 30, Tokyo Instrument Inc.). Fourier transform infrared (FT-IR) spectra were recorded

13

using an ABB Bomen MB100 Spectrometer (Bomen, Canada). Microscopic morphologies of the

14

samples were analyzed by field emission scanning electron microscopy (FE-SEM, Hitachi S-7400,

15

Hitachi, Japan) and transmission electron microscopy (TEM, JEM-2200, JEOL, Japan). The

16

specific surface area was determined using multipoint Brunauer, Emmett and Teller (BET) analysis.

17

The room-temperature photoluminescence (PL) spectra of the samples were investigating with a

18

luminescence spectrometer (LS 55; Perkin-Elmer Inc., USA) with a Xe lamp.

19

Electrochemical measurements

20

The cyclic voltammograms (CV) were taken in an N2-saturated 0.01M potassium ferrocyanide

21

solution prepared with 1M KNO3, with the scanning rate 50 mV/s on a

22

Potentiostat/Galvanostat/EIS (WonTech, ZIVE, SPI Korea) that was equipped with three-electrode

23

system. A platinum electrode, 1.6 mm in diameter and coated with the as-prepared catalyst paste

24

was used as the working electrode. A standard calomel electrode was used as the reference 7

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 32 of 70

1

electrode and platinum wire was used as the counter electrode. Similarly, electrochemical

2

impedance spectra (EIS) Nyquist plots measurements were taken in the same system over the

3

frequency range of 0.001 to 100 kHz. Before measurement, the working electrode was prepared as

4

follows: 20 mg of the different as-prepared samples were added to 20 µL of a solution of Nafion (5

5

wt %) and propanol. Subsequently, ultrasonication was carried out for 1 h to obtain a homogeneous

6

paste. Then 5 µg of the paste was taken and dripped on the surface of the platinum electrode to

7

form a thin catalyst film on the electrode. After drying at 80 °C in an oven for 30 min, the working

8

electrode was used for experiments.

9

Photocatalytic activity test

10

The photocatalytic activities of the prepared composites were evaluated by degrading RhB (10 ppm)

11

and RB-5 (10 ppm) under natural sunlight irradiation. The amount of the as-prepared composite,

12

initial concentration and the volume of the mixed solution were the same in all experiments. During

13

photocatalytic degradation, 20 mg of the as-prepared composite was mixed with 30 ml of the

14

organic dyes solution in a 100-ml beaker. Prior to photocatalytic reaction the mixed solution was

15

stirred in the dark for 30 min to reach desorption-absorption equilibrium. At a given time interval, 1

16

ml of the solution was withdrawn, centrifuged and filtered out, and, absorbance was measured.

17

Then the concentration of the dyes was analyzed using an UV−vis spectrophotometer. Following

18

equation was used to account the degradation rate, Degradation Efficiency % = ( −

C )× CO

%

19

where, Co is the initial concentration and C is the concentration after “t” minutes of irradiation. In

20

the reusability experiment, the recovered photocatalyst was centrifuged, filtered and dried. The

21

dried composite was used for the next run after making up for the lost portion with fresh solution.

22

All of the experiments were conducted in September, 2015 during sunny days in Jeonju, a city of

23

South Korea. 8

ACS Paragon Plus Environment

Page 33 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

Results and discussion

2

Characterization of the TCN samples

3

Figure 2 shows the XRD patterns of the different samples. It is noticeable that the synthesized TiO2

4

NFs and the composites with different wt % ratios of g-C3N4 exhibit similar patterns. The XRD

5

patterns reveal that most of the diffraction peaks of pure TiO2 correspond to the anatase phase. The

6

diffraction peak of pure g-C3N4 appearing at 27.4° matches to the (0 0 2) plane, which corresponds

7

to the characteristic interplanar stacking peak of the aromatic system 30. The peaks of TiO2 at

8

around 25.3°, 37.8°, 38.58°, 37.8°, 47.9°, 53.89°, 55.06°, 62.7°, 68.7°, 70.3° and 75.0° were attributed

9

to the (1 0 1), (0 0 4), (1 1 2), (2 0 0), (1 0 5), (2 1 1), (2 0 4), (1 1 6), (2 2 0) and (2 1 5) lattice

10

planes of anatase TiO2 (JCPDS, 21-1272). The peaks around 27.5°, 41.2° and 56.7° were attributed

11

to the (1 1 0), (1 1 1) and (2 2 0) lattice planes of rutile TiO2 (JCPDS 75-1755) 31. The anatase (1 0

12

1) peak for composites is much broader than that of pristine TiO2 (Figure 2 inset). This broadening

13

of peaks suggests that the lattice structure of TiO2 is distorted due to the intense interaction between

14

TiO2 and g-C3N4. Moreover, the diffraction peak of TiO2 at the higher wt % of g-C3N4 becomes

15

weaker, and also shifted slightly to a smaller value diffraction angle, further showing that the g-

16

C3N4 affects the crystal growth of TiO2. Thus, such broadened, weakened and shifted peaks can be

17

attributed to the intense interaction between TiO2 and g-C3N4 32. It should be noted that no typical

18

peak of g-C3N4 appears in the TiO2/g-C3N4 composite. The reason can be recognized to the low

19

weight loading of g-C3N4.

20

Figure 3 shows the Raman spectra of pure TiO2 and TCN-5 in the range of 0-1200 cm-1. Here,

21

Raman spectra were used to confirm the formation of TiO2 anatase phase and the attachment of g-

22

C3N4 sheet with TiO2 NFs. The pure TiO2 shows peaks at wave numbers of about 142 (Eg), 386

23

(B1g), 518 (A1g +B1g) and 640 (Eg) which can be attributed to the typical anatase phase of TiO2. 9

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 34 of 70

1

These observed peaks were assigned to the Ti—O—Ti network structure of TiO2 33. When the NFs

2

were attached on g-C3N4 sheets, there is a considerable increase in the intensity of these four

3

characteristic peaks. The possible reason can be credited to the fact that the porous TiO2 NFs are

4

attached on the surface of g-C3N4 sheets with strong chemical interaction. Moreover, the possible

5

formation of new bonds on the surface of TCN-5 composite was responsible for the enhancement

6

of the intensity at 142, 386, 518, and 640 cm−1. In addition, reducing properties of graphitic carbon

7

in carbon nitride facilitated the formation of a bulk anatase phase after the calcinations of TCN-5,

8

which also supports the enhanced intensity

9

prepared composite shows the slight shifting of Eg towards a higher wave number. This shift is

34, 35, 36

. Compared with pristine anatase TiO2, the as-

37

10

another indication for the formation of new bonds or bond modifications

11

analysis confirms that the TiO2 NFs could in fact be attaching to the surface of g-C3N4 sheets with

12

strong interaction rather than existing purely as a mixture of the two.

13

Regarding the FT-IR spectrum of bare TiO2 shown in Figure 4, the broad-band at around 500 to 800

14

cm-1 is assigned to the Ti—O—Ti stretching vibration mode. Additionally, the peak at 3400 cm-1

15

correspond to the stretching vibration mode of O—H bonds of free water molecules and Ti—OH,

16

and N—H vibrations for NH4+ ions 38. For the spectrum of bare g-C3N4, three characteristic

17

absorption regions were revealed located around 3200 cm-1, 1200-1650 cm-1, and 809 cm-1. Several

18

bands that were found in the 1200—1650 cm-1 corresponds to the typical stretching mode of

19

aromatic CN heterocycles 39. The peaks around 1640, 1462 and 1407 cm-1 is attributable to the

20

stretching vibration modes of heptazine-derived repeating units. In addition, the peaks at 1316 and

21

1240 cm-1 are allocated to the characteristic stretching vibrations of C—N(—C)—C 40. Similarly,

22

the broad adsorption band centered at 3188 cm-1 for g-C3N4 initiates from the stretching vibration

23

of the N—H bond associated with the uncondensed amino group. The strong absorption at 809 cm-1

24

could be attributed to the s-triazine ring system 41. As expected, all of the main characteristic 10

ACS Paragon Plus Environment

. Thus, the Raman

Page 35 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

absorption peaks of TiO2 and g-C3N4 were found to coexist in the TCN-x. Moreover, compared

2

with the peak at 3350 cm-1 for pure TiO2, a slight shift of this peak to the higher wavelength value

3

was observed in TCN-x, suggesting an interfacial interaction between TiO2 and g-C3N4 (Figure 4

4

inset). Such an intense interfacial interaction between the two compounds would enhance the

5

catalytic activity 25.

6

The morphologies of the prepared samples are observed by FE-SEM. The g-C3N4 showed irregular

7

surface morphology with different shapes and sizes of particles. However, its morphology has

8

significantly changed into sheet like structure after sonication. The different morphology and the

9

constituting elements of g-C3N4 are given in Figure S1. It was found that TTIP/PVAc composite

10

electrospun NFs from single nozzle electrospinning process, are uniform with smooth surfaces

11

before calcinations (Figure 5a). It can be observed from the figure that the TiO2 NFs become porous

12

and remained as continuous structures with mostly uniform diameters after calcinations. However,

13

their average diameter was decreased slightly (Figure 5b). The formation mechanism of porous NFs

14

is given in Figure S2. Figure 5c displays the morphology of composite electrospun mat of

15

TTIP/PVAc/g-C3N4 sheet prepared by means of angled two-nozzle electrospinning process. Here,

16

the nanosize particles were electrospun with and attached to PVAc fibers, whereas microsized

17

sheets were electrospun along with a number of PVAc fibers and attached on the composite mat.

18

The image clearly reveals the presence of g-C3N4 sheets in the composite mats. Moreover, the g-

19

C3N4 sheets do not seem to have a significant effect on the smoothness and uniformity of the NFs

20

indicating good performance of the electrospinning conditions. The FE-SEM images of electrospun

21

composite mats given in Figure S3 further clearly confirm the attachment of g-C3N4 sheets and the

22

morphology of the NFs. It is perceptible on the FE-SEM image that the TiO2 NFs are affixed on the

23

g-C3N4 sheet after calcinations in a reasonably unvarying manner (Figure 5d). The attached NFs

24

have a length of several micrometers with porous structure, but the diameter varies around 40 nm to 11

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 36 of 70

1

200 nm. In contrast to the NPs, such NFs of significant length promote direct and sufficient contact

2

with g-C3N4 sheet, resulting in less agglomeration. Such less agglomerated and uniformly

3

distributed NFs attached on the g-C3N4 sheets enhanced the surface area and porosity-related

4

characteristics of the composite than pristine TiO2 NFs, hence, further improving the

5

photodegradation efficiency. Moreover, such a very thin sheet, having large aspect ratios, high

6

surface area and a stoichiometric N/C ratio, increase the transport of charges and reduce the

7

recombination probability of photoexcited charge carriers.

8

Corresponding TEM images of TiO2 and TCN composites are displayed in Figure 6. Figure 6b

9

shows homogeneous, long and narrow, porous TiO2 NFs attached to the g-C3N4 sheet with a narrow

10

distribution of diameters ~ 125 nm. Moreover, HR-TEM image of the TCN composite suggests

11

intimate contact of TiO2 NFs and the g-C3N4 sheets. As shown in Figure 6c, a clear three sets of

12

lattice fringes of the interplanar spacing of about 0.48 nm, 0.35 nm and 0.35 nm are observed in the

13

dark area, which are very close to the (0 0 2), (1 0 1) and (1 0 1) planes of TiO2, respectively. The

14

electron beam can pass more easily through the g-C3N4 due to its low atomic weight, so the light

15

area can be determined to be g-C3N4 where it is very difficult to observe the lattice fringe.

16

The XPS survey spectrum and high resolution spectrum of the pristine and composite samples were

17

carried out to further illuminate the surface composition and chemical interaction between the

18

elements. Figure 7a illustrates the XPS survey spectrum of g-C3N4, TiO2 and TCN-5. Obviously,

19

the g-C3N4 contains the photoelectron peaks of C and N elements, and TCN-5 composite contains

20

the photoelectron peaks of C, N, Ti and O elements. However, the pure TiO2 NFs not only contain

21

the photoelectron peaks of Ti and O elements, but also contains the peak of C elements. This carbon

22

peak is ascribed to the residual carbon from the samples and adventitious hydrocarbon from the

23

XPS instrument itself. The deconvoluted peaks of all elements of the composite material are given 12

ACS Paragon Plus Environment

Page 37 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

in the corresponding high-resolution spectrum (Figure 7b-7e). Correspondingly, the high-resolution

2

XPS spectrum of all elements of the pristine TiO2 and g-C3N4 are given in Figure S4. The C 1s peak

3

of composite can be deconvoluted two fitted peaks at 285.65 eV and 287.05 eV (Figure 7b)

4

indicating that carbon posses two diverse chemical states 42. The peak around 285.65 eV is

5

attributed to defects in the g-C3N4 that involves sp2-hybridized carbon atoms. In addition, the other

6

peak located around 287.05 eV is assigned to N—C━N coordination 43, 44. The curves of N 1s

7

region can be divided also into two peaks situated at 397.7eV and 399.8 eV (Figure 7c). The main

8

peak located at 397.7eV belong to sp2 –hybridized pyridinic-like nitrogen(N-sp2C) and the peak

9

located at 399.8 eV corresponds to the tertiary pyrrolic graphitic nitrogen (N—(C)3) 45. The Ti 2p3/2

10

spin-orbital splitting photoelectron of the composite was located at binding energy 459.5 eV in the

11

Ti 2p spectrum. Similarly, peaks at 465.05 eV in the same spectrum correspond to the Ti 2p1/2

12

(Figure 7d) 46. The peak situated at 530.65 eV in the O 1s is ascribable to oxygen anions in the

13

lattice (Ti-O) (Figure 7e) 47. Furthermore, the binding energy values of Ti 2p in the composite are

14

slightly higher than those of pristine TiO2 (Figure 7f). Correspondingly, binding energy value of O

15

1s is also slightly higher in the composite compared to pristine TiO2 (Figure S5). Such shifts for

16

binding energy in the XPS spectra reflect the intense interaction between TiO2 and g-C3N4 48. Hence,

17

the results from XRD, Raman, FT-IR, TEM and XPS analyses strongly verified that the TiO2 NFs

18

are attached on the surface of g-C3N4 sheets forming a hetero-structure rather than a physical

19

mixture.

20

Figure 8a shows the UV-vis absorption spectra of different samples. The absorption intensity of the

21

pure g-C3N4 occurs at wavelengths greater than 450 nm, in good accordance with the band gap of

22

g-C3N4 (~ 2.7eV). However, the absorption wavelength of synthesized TiO2 NFs is less than 400

23

nm, consistent with the intrinsic band gap of anatase TiO2 (~ 3.18 eV) which means that it could

24

only have a response in the UV zone. The absorption edge of TCN-x composites shifts extremely to 13

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 38 of 70

1

a longer wavelength in comparison to the pure TiO2 NFs, clearly illuminating that the visible light

2

absorption was enhanced, which was because of the existence of g-C3N4 sheets. All of the TCN-x

3

composites exhibited absorbance in the visible light region, with the absorbance edges between

4

those of pure TiO2 and g-C3N4. It is noteworthy that the unattached TiO2 NFs express absorption in

5

the UV light region, while the TiO2 NFs attached on g-C3N4 sheets demonstrate absorption in the

6

visible light region. The band-gap energy of a semiconductor can be determined by the following

7

equation 49, 50, 51,

�ℎ� 8

1⁄

= � ℎ� − ��

where A is a proportional constant, ℎ� is the photo energy, h is Planck’s constant, � is the

10

frequency of vibration, � is an absorption coefficient and �� is the band gap energy. The value of

11

� depends on the transition in semiconductors. For g-C3N4 and TiO2, value of � is 1/2 for direct transition and 2 for indirect transition. Thus from the Figure 8b, the band gap energies of g-C3N4,

12

TiO2, TCN-1, TCN-3, TCN-5 and TCN-10 are calculated to be 3.35, 3.28, 3.0, 2.94, 2.8 and 2.73

13

eV respectively. The following empirical equations are used to calculate the CB and VB of

14

semiconductors 47,

9

�� = � − � + .5�� � 15 16 17 18

where �� is the valence band potential, �

= �� − �� is the conduction band potential, � is the electro

negativity which is calculated from the geometric mean of the constitute elements, � is the

energy of free electrons on the hydrogen scale which is 4.5 eV vs NHE, and �� is the band gap

energy. The � values are 5.81 and 4.64 for TiO2 and g-C3N4, respectively. From the above 14

ACS Paragon Plus Environment

Page 39 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

2

equations, �� and �

3

Correspondingly, �� and � respectively.

4

Photocatalytic activity

5

On the basis of the above results, photocatalytic activities of the different samples were assessed by

6

decomposing RhB and RB-5 under natural sunlight. Photodegradation was measured at 10 min and

7

5 min intervals for RhB and RB-5 respectively. Then the concentrations of RhB and RB-5 in the

8

solution were plotted as a function of irradiation time. The results of different catalysts are

9

displayed in Figure 9. As displayed in the figure, all TCN-x samples exhibited a better

1

values of g-C3N4 are calculated to be 1.5 and -1.23 eV, respectively. values of TiO2 are calculated to be 2.985 and -0.365 eV,

10

photocatalytic activity than pure TiO2 and g-C3N4 over RhB (Figure 9a) and RB-5 (Figure 9c),

11

suggesting that a synergistic effect exits between g-C3N4 and TiO2. This superior performance of

12

TCN-x over RhB and RB-5 could also be confirmed from the photographs shown in Figure 9b and

13

9d, respectively. In the absence of a photocatalyst, the self-degradation of both RhB and RB-5 is

14

negligible. Moreover, the figures show that the photocatalytic activity is enhanced with the content

15

of g-C3N4 increasing from 1 to 5 wt % of polymer weight. However, further increasing the g-C3N4

16

weight leads to a decrease in the degradation rate under the same photolysis condition. It is

17

noteworthy that the excess amount of carbon nitride in the composite creates an unsuitable ratio

18

between the TiO2 and g-C3N4, which may block the electron transfer from g-C3N4 sheets to TiO2

19

NFs and reduce the photogenerated charge separation. Thus, there is an optimum g-C3N4 content

20

for maximizing the photocatalytic activity. The TCN-5 samples demonstrate the highest

21

photocatalytic activity, which can absorb all most all of RhB and RB-5 after degradation for a short

22

period, 40 min and 20 min, respectively.

15

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 40 of 70

1

BET surface area of the samples was determined by a multipoint Brunauer-Emmett-Teller (BET)

2

method using nitrogen adsorption-desorption analysis. Table S1 summarizes the specific surface

3

area of pristine TiO2 and g-C3N4, and TCN-x composites. As shown in Table S1, the BET surface

4

area of TCN-x composites was higher than that of pristine TiO2 and g-C3N4. The reason behind this

5

is that the large surface area of g-C3N4 sheets provides ample space for TiO2 NFs, preventing the

6

agglomerating of NFs and increasing porosity of the composite, effectively increasing the surface

7

area. As can be seen from the table, the BET surface area of the composite gradually increases with

8

increasing amounts of g-C3N4 and showed the highest surface area (95.6313 m2/h) at TCN-5.

9

However, the BET surface area of TCN-10 was smaller than other composites indicating that this

10

may be an unsuitable ratio of TiO2 and g-C3N4. Thus, the higher surface area obtained in TCN-5

11

may be the result of a suitable ratio between TiO2 and g-C3N4. Correspondingly, the porous

12

characteristics, such as porosity, pore volume, pore size distribution and density of g-C3N4, TiO2

13

and TCN-5 samples were analyzed by mercury porosimetry analysis technique (Autopore IV,

14

Micromeritics USA). The different porous characteristics and distribution of pore diameters were

15

given in Table S2 and Figure S6, respectively. As shown in Figure S6, distribution of pore diameter

16

for TCN-5 and TiO2 were almost similar. However, the g-C3N4 showed lower pore size diameter

17

distribution compared to them. Here, the size of the pores means the space between two NFs. As

18

expected, porosity of TCN-5 was slightly higher than TiO2, which is in agreement with the FE-SEM

19

morphology. A possible reason for this is the large surface area of g-C3N4 sheets providing

20

sufficient space for the uniform attachment of TiO2 NFs which increased the porosity of the

21

composite. Hence, the TCN-5 composite, which has higher surface area and porosity, increased the

22

contact chance between pollutants and photocalayst, which is beneficial for improving the

23

photocatalytic degradation activity of the hybrids. However, the surface area and porosity are not

24

the only factors that affect the photoactivity. Other factors such as effective wavelength range and 16

ACS Paragon Plus Environment

Page 41 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

rate of recombination of e-h pairs also affect the photocatalytic activity 52. Hence, the TiO2 NFs

2

which have a higher surface area than g-C3N4, showed lower photocatalytic activity under sunlight

3

irradiation.

4

The Langmuir-Hinshelwood model is mainly used to estimate the kinetics of the photocatalytic

5

degradation

6

and degradation rate (r), which is expressed as 54,

53

. Their model basically narrates the concentration of reactants in water at time t(C)

�=−



=

�� �� � + �� �

7

where, Kr is the rate constant and Kad is the adsorption equilibrium constant. When the adsorption is

8

relatively weak and the concentration of reactants in a photocatalytic reaction is low, the

9

photocatalytic reaction equation can be simplified to a pseudo-first-order equation with an apparent

10

first-order rate constant (Kapp) as, �� (

� ) = �� �� � = �� � �

11

where, Co is the initial concentration of the substrate. Figure 10 shows the linear relationship

12

between ln(Co/C) and the irradiation time with Kapp for different catalyst. The Kapp value commonly

13

provides an indication of the activity of the photocatalyst. The Kapp value shows that the

14

photocatalytic activity for all TCN-x is higher than those of the pristine TiO2 and g-C3N4 for both

15

RhB and RB-5, which suggests that there exists a significant synergistic effect between g-C3N4 and

16

TiO2.

17

Figure 11 reveals the PL spectra of different samples in the wavelength range of 300-600 nm. The

18

first peak of TiO2 in the PL spectra around 370 nm is due to the direct recombination between

19

electrons in the conduction band and holes in the valence band 55. Additionally, the PL spectrum

20

peak around 480 nm and 530 nm is attributed to the surface oxygen vacancies and related defects 44.

21

Photo-induced electrons are attached by these oxygen vacancies and defects to form excitons, so 17

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 42 of 70

1

that PL signal can occur easily 56. Similarly, pure g-C3N4 exhibits a high intensity emission peak at

2

around 450 nm, which can be ascribed to the band-band PL phenomenon with a light energy that is

3

approximately equal to the band gap energy of g-C3N4 57. This band-band PL signal is a result of the

4

n →π* electronic transition involving lone pairs of nitrogen atoms in g-C3N4. Here, the all PL

5

intensity peaks of TCN-5 are lower in comparison to that of the pristine TiO2 and g-C3N4. Such a

6

lower PL intensity is a general indication of a lower recombination rate of e-h pairs 58.

7

In water-splitting photocatalysis, two processes are involved; oxidation process in which the holes

8

transfer from the photocatalyst to the degradable chemical in solution and a reduction process in

9

which electrons transfer from the photocatalyst to the solution. The better conductivity of g-C3N4

10

offers it as a super charge-carrier transport medium, and its large surface area may also bring about

11

a higher carrier transfer rate between the photocatalyst and the solution/photodegradable chemicals.

12

For better understanding of such a catalytic performance, a potentiostat/galvanostat/EIS analyzer

13

was used to measure the CV to confirm the interfacial charge transfer effect of TiO2, g-C3N4 and

14

TCN-5. Figure 12 reveals CV of different samples, in which clearer reduction and oxidation peaks

15

are observed for TCN-5 than other two pristine samples. As shown in the figure, the oxidation

16

peaks for g-C3N4, TiO2 and TCN-5 are located at 0.45, 0.57and 0.27 V and respectively represent

17

the oxidation of ferrocyanide with the loss of one electron. The peak current of TCN-5 (0.127

18

mA/cm2) is much higher than that of g-C3N4 (0.0505 mA/cm2) and TiO2 (0.039 mA/cm2)-more than

19

2.5 times, signifying a considerably improved rate of electron transfer due to g-C3N4 being a highly

20

conducting substrate. Furthermore, the ratio of the strengths of the oxidation (0.127 mA/cm2, 0.29V)

21

and reduction peaks (0.112 mA/cm2, 0.15V) of TCN-5 is nearly 1, which specifying greatly

22

enhanced reaction reversibility. EIS Nyquist plots were also recorded to further investigate the

23

interfacial charge immigration of the samples (Figure 13). It has been reported in recent research

24

that a smaller arc radius in Nyquist plots are related to a more effective interfacial charge 18

ACS Paragon Plus Environment

Page 43 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

immigration in semiconductor based electrode 59. Arc radius for TCN-5 in the Nyquist plot shown

2

in figure is smaller compared to g-C3N4 and TiO2, demonstrating the reduced interface resistance so

3

that the separation and migration efficiency of e-h pairs will be enhanced. Thus, combine with the

4

PL spectra, CV measurements and EIS Nyquist plots, it is concluded that the g-C3N4 sheet modified

5

by porous TiO2 NFs exhibits the most charge-carrier migration compared to ones.

6

A schematic diagram for the photocatalytic mechanism of TCN hetero-junction was proposed based

7

on the above experimental results and discussion (Figure 14). It was well understood that the

8

enrichment of the photodegradation activities of the composite materials was mainly to be realized

9

via the effective transfer of electron and holes at the interface of the photocatalysts. When TCN-x

10

was exposed to the visible light, g-C3N4 could easily absorb the visible light because of its band gap

11

(2.73 eV), leading to the excitation of electrons and generation of e-h pairs. Given the calculated

12

value above, the CB edge potential of g-C3N4 (-1.23 eV) is more negative than that CB of TiO2

13

(-0.365 eV). Hence the photogenerated electrons transfer easily from the CB of g-C3N4 to the CB of

14

TiO2. It has been reported in various previous research that the difference in the CB edge potential

15

between these two materials were perhaps a more powerful driving force, which promote the

16

transfer of electrons between hetero-junctions and reduce the recombination of charge carriers 58, 60.

17

Since the CB edge potentials of g-C3N4 and TiO2 were more negative than the reduction potential of

18

oxygen E° (O2/.O2̶ ) (-0.046 eV) 61, electrons transferred to the CB of TiO2 NFs,

19

main active groups of superoxide anions radicals (.O2̶) through reacting with dissolved oxygen near

20

the surface of TiO2, which is mainly responsible for the degradation of the pollutants. In addition,

21

since the VB edge potential of TiO2 (2.985 eV) is more positive than the VB of g-C3N4 (1.5 eV),

22

transfer of photogenerated holes occurs from the VB of TiO2 to the VB of g-C3N4 through the close

23

interfacial connection between g-C3N4 and TiO2 61, 62. However, the hole on the VB of g-C3N4 could

24

not generate .OH radicals through reacting with OH− or H2O because of the lower VB level of g19

ACS Paragon Plus Environment

generating the

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 44 of 70

1

C3N4 compared to the potential of .OH/OH− (1.99 eV) 63. Conversely, the remaining holes on the

2

VB of TiO2 directly oxidize the pollutants and concurrently generate .OH radicals through reacting

3

with H2O because of the higher VB level of TiO2 compared to the potential of .OH/OH−. Since

4

most of the holes are transferred from the VB of TiO2 to the VB of g-C3N4, the number of holes that

5

remain in the VB of TiO2 is considerably reduced. Therefore, in TCN composites, the .O2̶ radicals

6

and electron are the most active species during photocatalysis compared to .OH and holes. However,

7

for pure g-C3N4, most of the photogenerated electrons and holes recombine due to the narrow

8

energy gap (2.73 eV), and only a small fraction participate in the photocatalytic reaction, resulting

9

in lower activity 27. Correspondingly, because of broader band gap (Eg = 3.35 eV), the anatase TiO2

10

could not be exited under 420 nm visible light irradiation, thus showing a weaker photocatalytic

11

degradation activity. In most of the previous studies of TiO2 and g-C3N4 composites, TiO2 NPs were

12

dispersed on the g-C3N4 sheet. Such NPs have fewer chances for uniform distribution over the

13

surface of the g-C3N4 sheet without agglomeration. As mentioned above, NFs have more uniform

14

distribution over the surface of g-C3N4 sheets so more NFs have direct attachment with the g-C3N4

15

sheet. Moreover, such porous structured NFs, having enhanced surface area and porosity, could

16

promote the migration of electrons. Therefore, electron transfer in such a straight path between g-

17

C3N4 sheet and NFs will be much easier in comparison to the zigzag path in agglomerated NPs.

18

The recycling capability of the TCN-5 photocatalyst was investigated, by degrading RhB under

19

visible light, cycled for four times. For this, the recovered photocatalyst was centrifuged, filtered,

20

and dried at 60 °C overnight. This dried composite was used for the next run after making up the

21

lost portion with fresh material. As displayed in Figure 15, the photocatalytic activity of TCN-5 for

22

RhB degradation decreased from 91.26% to 84.56% after the four-run test, which demonstrates that

23

it has satisfactory stability of the photocatalytic activity. Moreover, FT-IR patterns were used to

24

further investigate the lifetime of g-C3N4, TiO2 and TCN-5. For this, the photocatalysts after 4 20

ACS Paragon Plus Environment

Page 45 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

cyclic runs were centrifuged, filtered and dried at 60 °C overnight. Then, the FT-IR spectra of these

2

used materials were taken and compared with fresh one. Figure S7 clearly illustrates that the FT-IR

3

patterns of all samples after four cyclic runs had not changed depreciably compared to the fresh

4

sample. This result hints that the lifetime of these samples are satisfactory for photocatalytic

5

degradation activity.

6

Conclusions

7

In summary, the porous TiO2 NFs became well attached to the surface of the sheet-like structure of

8

g-C3N4 through the use of a simple and facile electrospinning-calcination process. After

9

incorporation of porous TiO2 NFs, the TiO2/g-C3N4 composites exhibited significantly enhanced

10

photocatalytic degradation over RhB and RB-5 than did pristine TiO2 and g-C3N4, under irradiation

11

by natural sunlight. The PL spectra, CV and EIS Nyquist plots measurements suggest that the better

12

conductivity of g-C3N4 provides a super charge-carrier transport medium and a higher carrier

13

transfer rate between the photocatalyst and the solution, by holding down the recombination of e-h

14

pairs in TiO2, leading to the extended lifetime of charge carriers and improved photocatalytic

15

activity.

16

Associated Content

17

Supporting Information

18

FE-SEM images, SEM-EDX image with real atomic compositions and XPS survey spectrum of g-

19

C3N4; porous formation mechanism of TiO2 NFs; FE-SEM images of electrospun PVAc NFs with

20

g-C3N4 sheets in different magnifications; high resolution XPS spectra of Ti 2p, O1s, C1s, and N1s

21

for the pristine g-C3N4 and TiO2 NFs; binding energy shifting comparison of O1s high resolution

22

XPS spectra from pristine TiO2 NFs and TCN-5 composite; BET surface area and Langmuir surface

23

area of different samples; distribution of pore diameter.

24

21

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 46 of 70

1

Acknowledgement

2

This paper was supported by the Human Resource Training Program for Regional Innovation and

3

Creativity through the Ministry of Education and National Research Foundation of Korea (NRF-

4

2015H1C1A1035635)

5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

22

ACS Paragon Plus Environment

Page 47 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

References

2

1.

3

Activity of TiO2 Nanocrystals with Internal Pores. ACS Applied Materials & Interfaces 2014, 6 (3),

4

1608-1615.

5

2.

6

for UV-photodegradation of methylene blue. Applied Catalysis B: Environmental 2006, 63 (3–4),

7

305-312.

8

3.

9

a Visible Light Photocatalyst. The New Role of Graphene as a Macromolecular Photosensitizer.

Ren, L.; Li, Y.; Hou, J.; Zhao, X.; Pan, C. Preparation and Enhanced Photocatalytic

Height, M. J.; Pratsinis, S. E.; Mekasuwandumrong, O.; Praserthdam, P. Ag-ZnO catalysts

Zhang, Y.; Zhang, N.; Tang, Z.-R.; Xu, Y.-J. Graphene Transforms Wide Band Gap ZnS to

10

ACS Nano 2012, 6 (11), 9777-9789.

11

4.

12

Hollow Spheres with Enhanced Photocatalytic Activity and Durability. ACS Applied Materials &

13

Interfaces 2012, 4 (3), 1813-1821.

14

5.

15

WO3 semiconductor photocatalyst prepared from amorphous peroxo-tungstic acid for the

16

degradation of various organic compounds. Applied Catalysis B: Environmental 2010, 94 (1–2),

17

150-157.

18

6.

19

Microspheres with High Photocatalytic Activity. Crystal Growth & Design 2008, 8 (3), 930-934.

20

7.

21

CO2 with homogeneously dispersed nanoscale TiO2 catalysts. Chemical Communications 2004,

22

(10), 1234-1235.

23

8.

24

Society Reviews 2008, 37 (10), 2328-2353.

Luo, M.; Liu, Y.; Hu, J.; Liu, H.; Li, J. One-Pot Synthesis of CdS and Ni-Doped CdS

Sayama, K.; Hayashi, H.; Arai, T.; Yanagida, M.; Gunji, T.; Sugihara, H. Highly active

Yu, J.; Liu, W.; Yu, H. A One-Pot Approach to Hierarchically Nanoporous Titania Hollow

Pathak, P.; Meziani, M. J.; Li, Y.; Cureton, L. T.; Sun, Y.-P. Improving photoreduction of

Lun Pang, C.; Lindsay, R.; Thornton, G. Chemical reactions on rutile TiO2(110). Chemical

23

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 48 of 70

1

9.

Hou, H.; Wang, L.; Gao, F.; Wei, G.; Zheng, J.; Tang, B.; Yang, W. Fabrication of porous

2

titanium dioxide fibers and their photocatalytic activity for hydrogen evolution. International

3

Journal of Hydrogen Energy 2014, 39 (13), 6837-6844.

4

10.

5

and Electrospun TiO2 Nanofibers: Effects of Mesoporosity and Interparticle Charge Transfer. The

6

Journal of Physical Chemistry C 2010, 114 (39), 16475-16480.

7

11.

8

Heterojunctions in g-C3N4/TiO2(B) nanofibres with exposed (001) plane and enhanced visible-light

9

photoactivity. Journal of Materials Chemistry A 2014, 2 (7), 2071-2078.

Choi, S. K.; Kim, S.; Lim, S. K.; Park, H. Photocatalytic Comparison of TiO2 Nanoparticles

Zhang, L.; Jing, D.; She, X.; Liu, H.; Yang, D.; Lu, Y.; Li, J.; Zheng, Z.; Guo, L.

10

12.

Liu, Z.; Sun, D. D.; Guo, P.; Leckie, J. O. An Efficient Bicomponent TiO2/SnO2 Nanofiber

11

Photocatalyst Fabricated by Electrospinning with a Side-by-Side Dual Spinneret Method. Nano

12

Letters 2007, 7 (4), 1081-1085.

13

13.

14

SrTiO3/TiO2 Nanofiber Heterostructures with High Photocatalytic Activity. Langmuir 2011, 27 (6),

15

2946-2952.

16

14.

17

strategies towards visible light responsive photoanode, a review. Energy & Environmental Science

18

2011, 4 (4), 1065-1086.

19

15.

20

Properties, and Applications in Catalysis. ACS Applied Materials & Interfaces 2014, 6 (19), 16449-

21

16465.

22

16.

23

photocatalytic H2 evolution under visible light. Chemical Communications 2012, 48 (28), 3430-

24

3432.

Cao, T.; Li, Y.; Wang, C.; Shao, C.; Liu, Y. A Facile in Situ Hydrothermal Method to

El Ruby Mohamed, A.; Rohani, S. Modified TiO2 nanotube arrays (TNTAs): progressive

Zhu, J.; Xiao, P.; Li, H.; Carabineiro, S. A. C. Graphitic Carbon Nitride: Synthesis,

Yan, H. Soft-templating synthesis of mesoporous graphitic carbon nitride with enhanced

24

ACS Paragon Plus Environment

Page 49 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

17.

Kailasam, K.; Epping, J. D.; Thomas, A.; Losse, S.; Junge, H. Mesoporous carbon nitride-

2

silica composites by a combined sol-gel/thermal condensation approach and their application as

3

photocatalysts. Energy & Environmental Science 2011, 4 (11), 4668-4674.

4

18.

5

over Boron-Doped g-C3N4 under Visible Light Irradiation. Langmuir 2010, 26 (6), 3894-3901.

6

19.

7

Compounds: A New Functional Organic–Metal Hybrid Material. Advanced Materials 2009, 21 (16),

8

1609-1612.

9

20.

Yan, S. C.; Li, Z. S.; Zou, Z. G. Photodegradation of Rhodamine B and Methyl Orange

Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-Containing Carbon Nitride

Ge, L.; Han, C. Synthesis of MWNTs/g-C3N4 composite photocatalysts with efficient

10

visible light photocatalytic hydrogen evolution activity. Applied Catalysis B: Environmental 2012,

11

117–118 (0), 268-274.

12

21.

13

Sulfur Flux for Water Photoredox Catalysis. ACS Catalysis 2012, 2 (6), 940-948.

14

22.

15

of ZnO hybridized with graphite-like C3N4. Energy & Environmental Science 2011, 4 (8), 2922-

16

2929.

17

23.

18

synthesis, enhanced activity and photocatalytic mechanism. Journal of Materials Chemistry 2012,

19

22 (39), 21159-21166.

20

24.

21

various TiO2 nanostructures on g-C3N4 nanosheets with much enhanced photocatalytic activity

22

under visible light. J. Hazard. Mater. 2015, 292, 79-89.

23

25.

24

scheme g-C3N4-TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem.

Zhang, J.; Zhang, M.; Zhang, G.; Wang, X. Synthesis of Carbon Nitride Semiconductors in

Wang, Y.; Shi, R.; Lin, J.; Zhu, Y. Enhancement of photocurrent and photocatalytic activity

Fu, J.; Tian, Y.; Chang, B.; Xi, F.; Dong, X. BiOBr-carbon nitride heterojunctions:

Li, Y.; Wang, J.; Yang, Y.; Zhang, Y.; He, D.; An, Q.; Cao, G. Seed-induced growing

Yu, J.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-

25

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 50 of 70

1

Chem. Phys. 2013, 15 (39), 16883-16890.

2

26.

3

composite nanofibers by electrospinning for increased photocatalysis. Carbon 2012, 50 (7), 2472-

4

2481.

5

27.

6

Improved Photocatalytic Activities. Advanced Functional Materials 2012, 22 (22), 4763-4770.

7

28.

8

TiO2 Nanowire and Graphene-TiO2 Nanoparticle Composite Photocatalysts. ACS Applied Materials

9

& Interfaces 2012, 4 (8), 3944-3950.

Kim, C. H.; Kim, B.-H.; Yang, K. S. TiO2 nanoparticles loaded on graphene/carbon

Niu, P.; Zhang, L.; Liu, G.; Cheng, H.-M. Graphene-Like Carbon Nitride Nanosheets for

Pan, X.; Zhao, Y.; Liu, S.; Korzeniewski, C. L.; Wang, S.; Fan, Z. Comparing Graphene-

10

29.

Han, C.; Wang, Y.; Lei, Y.; Wang, B.; Wu, N.; Shi, Q.; Li, Q. In situ synthesis of graphitic-

11

C3N4 nanosheet hybridized N-doped TiO2 nanofibers for efficient photocatalytic H2 production and

12

degradation. Nano Res. 2015, 8 (4), 1199-1209.

13

30.

14

ZnO flowers on the surface of g-C3N4 sheets via hydrothermal process. Ceramics International

15

2015, 41 (10, Part A), 12923-12929.

16

31.

17

Anatase TiO2 Nanoparticles on Rutile TiO2 Nanorods:  A Heterogeneous Nanostructure via Layer-

18

by-Layer Assembly. Langmuir 2007, 23 (22), 10916-10919.

19

32.

20

Hydrothermal Synthesis of Graphene-TiO2 Nanotube Composites with Enhanced Photocatalytic

21

Activity. ACS Catalysis 2012, 2 (6), 949-956.

22

33.

23

ball milling method. Phys. Chem. Chem. Phys. 2015, 17 (5), 3647-3652.

24

34.

Prasad Adhikari, S.; Raj Pant, H.; Joo Kim, H.; Hee Park, C.; Sang Kim, C. Deposition of

Liu, Z.; Zhang, X.; Nishimoto, S.; Jin, M.; Tryk, D. A.; Murakami, T.; Fujishima, A.

Perera, S. D.; Mariano, R. G.; Vu, K.; Nour, N.; Seitz, O.; Chabal, Y.; Balkus, K. J.

Zhou, J.; Zhang, M.; Zhu, Y. Photocatalytic enhancement of hybrid C 3N4/TiO2 prepared via

Song, P.; Zhang, X.; Sun, M.; Cui, X.; Lin, Y. Graphene oxide modified TiO2 nanotube 26

ACS Paragon Plus Environment

Page 51 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

arrays: enhanced visible light photoelectrochemical properties. Nanoscale 2012, 4 (5), 1800-1804.

2

35.

3

C3N4/B-TiO2 nanosheets with exposed {001} plane and enhanced visible-light photocatalytic

4

activities. International Journal of Hydrogen Energy 2016, 41 (18), 7292-7300.

5

36.

6

supported ultralong TiO2 nanofibers from the commercial titania for high performance lithium-ion

7

batteries. Journal of Materials Chemistry A 2015, 3 (12), 6642-6648.

8

37.

9

Doping on the Visible-Light Photocatalytic Activity of Mesoporous TiO2. Angewandte Chemie

Chen, L.; Zhou, X.; Jin, B.; Luo, J.; Xu, X.; Zhang, L.; Hong, Y. Heterojunctions in g-

Gu, G.; Cheng, J.; Li, X.; Ni, W.; Guan, Q.; Qu, G.; Wang, B. Facile synthesis of graphene

Liu, G.; Zhao, Y.; Sun, C.; Li, F.; Lu, G. Q.; Cheng, H.-M. Synergistic Effects of B/N

10

International Edition 2008, 47 (24), 4516-4520.

11

38.

12

Temperature on the Surface Microstructure and Photocatalytic Activity of TiO2 Thin Films Prepared

13

by Liquid Phase Deposition. The Journal of Physical Chemistry B 2003, 107 (50), 13871-13879.

14

39.

15

Directly Heating Melamine. Langmuir 2009, 25 (17), 10397-10401.

16

40.

17

carbon nitride with recyclable adsorption and photocatalytic activity. Journal of Materials

18

Chemistry 2011, 21 (38), 14398-14401.

19

41.

20

Pt-free hydrogen production. The co-catalyst function of Ag/Ag2S formed by simultaneous

21

photodeposition. Dalton Transactions 2014, 43 (12), 4878-4885.

22

42.

23

like anatase TiO2 thin film for photocatalytic oxidation of bisphenol A. Applied Catalysis B:

24

Environmental 2011, 110, 260-272.

Yu, J.-G.; Yu, H.-G.; Cheng, B.; Zhao, X.-J.; Yu, J. C.; Ho, W.-K. The Effect of Calcination

Yan, S. C.; Li, Z. S.; Zou, Z. G. Photodegradation Performance of g-C3N4 Fabricated by

Liu, J.; Zhang, T.; Wang, Z.; Dawson, G.; Chen, W. Simple pyrolysis of urea into graphitic

Jiang, D.; Chen, L.; Xie, J.; Chen, M. Ag2S/g-C3N4 composite photocatalysts for efficient

Ng, J.; Wang, X.; Sun, D. D. One-pot hydrothermal synthesis of a hierarchical nanofungus-

27

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 52 of 70

1

43.

Liu, B.; Zeng, H. C. Carbon Nanotubes Supported Mesoporous Mesocrystals of Anatase

2

TiO2. Chemistry of Materials 2008, 20 (8), 2711-2718.

3

44.

4

SWCNTs (C60-d-CNTs) and its TiO2-based nanocomposite with enhanced photocatalytic activity

5

for hydrogen production. Dalton Transactions 2013, 42 (10), 3402-3409.

6

45.

7

TiO2 and g-C3N4 on the photoreactivity of g-C3N4/TiO2 photocatalyst: (0 0 1) vs (1 0 1) facets of

8

TiO2. Applied Catalysis B: Environmental 2015, 164, 420-427.

9

46.

Chai, B.; Peng, T.; Zhang, X.; Mao, J.; Li, K.; Zhang, X. Synthesis of C60-decorated

Huang, Z. a.; Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of contact interface between

Fu, M.; Liao, J.; Dong, F.; Li, H.; Liu, H. Growth of g-C3N4 Layer on Commercial TiO2 for

10

Enhanced Visible Light Photocatalytic Activity. Journal of Nanomaterials 2014, 2014, 8.

11

47.

12

C3N4/Ag/TiO2 Microspheres with Enhanced Photocatalysis Performance under Visible-Light

13

Irradiation. ACS Applied Materials & Interfaces 2014, 6 (16), 14405-14414.

14

48.

15

coupled with g-C3N4 nanosheets as 2D/2D heterojunction photocatalysts toward high photocatalytic

16

activity. Applied Catalysis B: Environmental 2015, 163, 298-305.

17

49.

18

Chen, J.; Inceesungvorn, B. Enhanced visible-light photocatalytic activity of g-C3N4/TiO2 films. J.

19

Colloid Interface Sci. 2014, 417, 402-409.

20

50.

21

Synthesis and photocatalytic activities of CdS/TiO2 nanoparticles supported on carbon nanofibers

22

for high efficient adsorption and simultaneous decomposition of organic dyes. J. Colloid Interface

23

Sci. 2014, 434, 159-166.

24

51.

Chen, Y.; Huang, W.; He, D.; Situ, Y.; Huang, H. Construction of Heterostructured g-

Zhang, Z.; Huang, J.; Zhang, M.; Yuan, Q.; Dong, B. Ultrathin hexagonal SnS 2 nanosheets

Boonprakob, N.; Wetchakun, N.; Phanichphant, S.; Waxler, D.; Sherrell, P.; Nattestad, A.;

Pant, B.; Barakat, N. A. M.; Pant, H. R.; Park, M.; Saud, P. S.; Kim, J.-W.; Kim, H.-Y.

Barakat, N. A. M.; Khil, M. S.; Sheikh, F. A.; Kim, H. Y. Synthesis and Optical Properties 28

ACS Paragon Plus Environment

Page 53 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1

of Two Cobalt Oxides (CoO and Co3O4) Nanofibers Produced by Electrospinning Process. The

2

Journal of Physical Chemistry C 2008, 112 (32), 12225-12233.

3

52.

4

and N-doped TiO2 synthesised at different thermal treatment temperatures with the same

5

hydrothermal precursor. Dalton Transactions 2014, 43 (36), 13783-13791.

6

53.

7

Photocatalytic degradation of phenol in natural seawater using visible light active carbon modified

8

(CM)-n-TiO2 nanoparticles under UV light and natural sunlight illuminations. Chemosphere 2013,

9

91 (3), 307-313.

Wang, J.; Fan, C.; Ren, Z.; Fu, X.; Qian, G.; Wang, Z. N-doped TiO2/C nanocomposites

Shaban, Y. A.; El Sayed, M. A.; El Maradny, A. A.; Al Farawati, R. K.; Al Zobidi, M. I.

10

54.

Kartal, O. E.; Erol, M.; Oguz, H. Photocatalytic destruction of phenol by TiO2 powders.

11

Chem. Eng. Technol. 2001, 24 (6), 645-649.

12

55.

13

and characterization of La doped TiO2 nanoparticles and their photocatalytic activity. Journal of

14

Solid State Chemistry 2004, 177 (10), 3375-3382.

15

56.

16

and the Photocatalytic Property of ZnO Nanocrystals in Nafion Membranes. Langmuir 2009, 25 (2),

17

1218-1223.

18

57.

19

high visible light photoelectrochemical activity. Electrochimica Acta 2012, 71, 10-16.

20

58.

21

nanotube array heterojunction visible-light photocatalyst: synthesis, characterization, and

22

photoelectrochemical properties. Journal of Materials Chemistry 2012, 22 (34), 17900-17905.

23

59.

24

enhanced photoactivity of water oxidation. Energy & Environmental Science 2011, 4 (5), 1781-

Liqiang, J.; Xiaojun, S.; Baifu, X.; Baiqi, W.; Weimin, C.; Honggang, F. The preparation

Wang, J.; Liu, P.; Fu, X.; Li, Z.; Han, W.; Wang, X. Relationship between Oxygen Defects

Wang, J.; Zhang, W.-D. Modification of TiO2 nanorod arrays by graphite-like C3N4 with

Zhou, X.; Jin, B.; Li, L.; Peng, F.; Wang, H.; Yu, H.; Fang, Y. A carbon nitride/TiO2

Hong, S. J.; Lee, S.; Jang, J. S.; Lee, J. S. Heterojunction BiVO4/WO3 electrodes for

29

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 54 of 70

1

1787.

2

60.

3

visible-light-driven photocatalyst for As3+ oxidation, MO degradation and water splitting for

4

hydrogen evolution. Journal of Materials Chemistry A 2014, 2 (38), 15774-15780.

5

61.

6

nanocomposite as an efficient photocatalyst for hydrogen production under visible light irradiation.

7

Phys. Chem. Chem. Phys. 2012, 14 (48), 16745-16752.

8

62.

9

C3N4 and TaON with improved visible light photocatalytic activities. Dalton Transactions 2010, 39

Zang, Y.; Li, L.; Xu, Y.; Zuo, Y.; Li, G. Hybridization of brookite TiO2 with g-C3N4: a

Chai, B.; Peng, T.; Mao, J.; Li, K.; Zan, L. Graphitic carbon nitride (g-C3N4)-Pt-TiO2

Yan, S. C.; Lv, S. B.; Li, Z. S.; Zou, Z. G. Organic-inorganic composite photocatalyst of g-

10

(6), 1488-1491.

11

63.

12

Synthesis of Ti3+ Self-Doped TiO2/g-C3N4 Heterojunctions with High Photocatalytic Performance

13

under LED Light Irradiation. ACS Applied Materials & Interfaces 2015, 7 (17), 9023-9030.

Li, K.; Gao, S.; Wang, Q.; Xu, H.; Wang, Z.; Huang, B.; Dai, Y.; Lu, J. In-Situ-Reduced

14 15 16 17 18 19 20 21 22

30

ACS Paragon Plus Environment

Page 55 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 1: Schematic illustration for the synthesis of the TiO2/g-C3N4 composite.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

31

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 56 of 70

1 2

Figure 2: XRD patterns of (a) g-C3N4, (b) TiO2, (c) TCN-1, (d) TCN-3, (e) TCN-5, and (f) TCN-10.

3

Inset XRD pattern showing the 23-30 degree region.

4 5 6 7 8 9 10 11 12

32

ACS Paragon Plus Environment

Page 57 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 3: Raman shift of TiO2 and TCN-5.

3 4 5 6 7 8 9 10 11 12 13 14

33

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 58 of 70

1 2

Figure 4: FT-IR images of (a) g-C3N4, (b) TiO2, (c) TCN-1, (d) TCN-3, (e) TCN-5 and (f) TCN-10.

3

Inset FT-IR spectra showing 2500-4000 cm-1 region.

4 5 6 7 8 9 10 11

34

ACS Paragon Plus Environment

Page 59 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 5: FE-SEM images of (a) TTIP/PVAC fibers before calcinations, (b) TiO2 fibers after

3

calcination, (c) TTIP/PVAC/g-C3N4 fibers before calcinations, and (d) TCN composite after

4

calcinations.

5 6 7 8 9 10 11 12

35

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 60 of 70

1 2

Figure 6: (a) TEM images of (a) TiO2 NFs, (b) TCN, and (c) HR-TEM images of TCN composite.

3 4 5 6 7 8 9 10

36

ACS Paragon Plus Environment

Page 61 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 7: (a) XPS survey spectrum and high resolution XPS spectra of (b) C1s, (c) N1s, (d) Ti 2p,

3

(e) O1s and (f) shifting of Ti2p peak in TCN-5 composite.

4 5

37

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 62 of 70

1 2

Figure 8: (a) UV-vis absorption spectra of TiO2, g-C3N4 and TCN-x samples and (b) plots of (αhν)2

3

vs. photon energy (hν) for the band gap energies of different samples.

4 5 6 7 8 9 10 11

38

ACS Paragon Plus Environment

Page 63 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 9: (a) Photodegradation of RhB under natural sun light in the corresponding time intervals;

3

(b) Photo images showing the initial 10-ppm RhB solution and the remaining RhB in the solution

4

after being irradiated for 40 min; (c) Photodegradation of RB-5 under natural sunlight in the

5

corresponding time intervals; and (d) Photo images showing the initial 10- ppm RB-5 solution and

6

the remaining RB-5 in the solution after being irradiated for 20 min.

7 8 9 10 11 12 13 14

39

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 64 of 70

1 2

Figure 10: Kinetics of photocatalytic degradation of different catalysts, (a) 10-ppm RhB solution,

3

and (b) 10-ppm RB-5 solution.

4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

40

ACS Paragon Plus Environment

Page 65 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 11: PL spectra of g-C3N4, TiO2 and TCN-5

3 4 5 6 7 8 9 10

41

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

1 2

Figure 12: Cyclic voltammograms of TiO2, g-C3N4 and TCN-5

3 4 5 6 7 8 9 10 11 12 13

42

ACS Paragon Plus Environment

Page 66 of 70

Page 67 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 13: Electrochemical impedance spectra of TiO2, g-C3N4 and TCN-5

3 4 5 6 7 8 9 10 11 12

43

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

1 2

Figure 14: Schematic diagram of photogenerated charges in the TCN hetero-junction.

3 4 5 6 7 8 9 10 11

44

ACS Paragon Plus Environment

Page 68 of 70

Page 69 of 70

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Langmuir

1 2

Figure 15: Cyclic runs of TCN-5 in the photocatalytic degradation over RhB under natural light

3

irradiation.

4 5 6 7 8 9 10 11 12 13 14 15 16 17

45

ACS Paragon Plus Environment

Langmuir

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 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

Page 70 of 70

1

Electrospinning directly synthesized porous TiO2 nanofibers modified by graphitic carbon

2

nitride sheets for enhanced photocatalytic degradation activity under solar light irradiation

3

Table of Contents Graphic

4

46

ACS Paragon Plus Environment