p-block-based ferroelectric-photocatalyst compounds

126
University of Wollongong University of Wollongong Research Online Research Online University of Wollongong Thesis Collection 2017+ University of Wollongong Thesis Collections 2018 P-block-based Ferroelectric-photocatalyst compounds: Structure, P-block-based Ferroelectric-photocatalyst compounds: Structure, Ferroelectric properties and photocatalytic performance Ferroelectric properties and photocatalytic performance Amar Hadee Jareeze Al-Keisy Follow this and additional works at: https://ro.uow.edu.au/theses1 University of Wollongong University of Wollongong Copyright Warning Copyright Warning You may print or download ONE copy of this document for the purpose of your own research or study. The University does not authorise you to copy, communicate or otherwise make available electronically to any other person any copyright material contained on this site. You are reminded of the following: This work is copyright. Apart from any use permitted under the Copyright Act 1968, no part of this work may be reproduced by any process, nor may any other exclusive right be exercised, without the permission of the author. Copyright owners are entitled to take legal action against persons who infringe their copyright. A reproduction of material that is protected by copyright may be a copyright infringement. A court may impose penalties and award damages in relation to offences and infringements relating to copyright material. Higher penalties may apply, and higher damages may be awarded, for offences and infringements involving the conversion of material into digital or electronic form. Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily represent the views of the University of Wollongong. represent the views of the University of Wollongong. Recommended Citation Recommended Citation Al-Keisy, Amar Hadee Jareeze, P-block-based Ferroelectric-photocatalyst compounds: Structure, Ferroelectric properties and photocatalytic performance, Doctor of Philosophy thesis, Institute for Superconducting and Electronic Materials, University of Wollongong, 2018. https://ro.uow.edu.au/ theses1/329 Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected]

Upload: others

Post on 01-Jan-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: P-block-based Ferroelectric-photocatalyst compounds

University of Wollongong University of Wollongong

Research Online Research Online

University of Wollongong Thesis Collection 2017+ University of Wollongong Thesis Collections

2018

P-block-based Ferroelectric-photocatalyst compounds: Structure, P-block-based Ferroelectric-photocatalyst compounds: Structure,

Ferroelectric properties and photocatalytic performance Ferroelectric properties and photocatalytic performance

Amar Hadee Jareeze Al-Keisy

Follow this and additional works at: https://ro.uow.edu.au/theses1

University of Wollongong University of Wollongong

Copyright Warning Copyright Warning

You may print or download ONE copy of this document for the purpose of your own research or study. The University

does not authorise you to copy, communicate or otherwise make available electronically to any other person any

copyright material contained on this site.

You are reminded of the following: This work is copyright. Apart from any use permitted under the Copyright Act

1968, no part of this work may be reproduced by any process, nor may any other exclusive right be exercised,

without the permission of the author. Copyright owners are entitled to take legal action against persons who infringe

their copyright. A reproduction of material that is protected by copyright may be a copyright infringement. A court

may impose penalties and award damages in relation to offences and infringements relating to copyright material.

Higher penalties may apply, and higher damages may be awarded, for offences and infringements involving the

conversion of material into digital or electronic form.

Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily

represent the views of the University of Wollongong. represent the views of the University of Wollongong.

Recommended Citation Recommended Citation Al-Keisy, Amar Hadee Jareeze, P-block-based Ferroelectric-photocatalyst compounds: Structure, Ferroelectric properties and photocatalytic performance, Doctor of Philosophy thesis, Institute for Superconducting and Electronic Materials, University of Wollongong, 2018. https://ro.uow.edu.au/theses1/329

Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected]

Page 2: P-block-based Ferroelectric-photocatalyst compounds

P-block-based Ferroelectric-photocatalyst compounds:

Structure, Ferroelectric properties and photocatalytic

performance

Amar Hadee Jareeze Al-Keisy: B.&M. physics

Supervisors: Dr. Yi Du & Prof. Shi Xue Dou

This thesis is presented as part of the requirement for the conferral of the degree: Doctor of Philosophy

The University of Wollongong

Institute for Superconducting and Electronic Materials

12/2018

Page 3: P-block-based Ferroelectric-photocatalyst compounds

i

Abstract

The high recombination rates of photogenerated electron-hole pairs inside the bulk of

photocatalyst, the back-reactions of intermediate species and inactive visible light are regarded as

the key issues blocking solar light-conversion efficiency. P-block photocatalyst based-internal

electric field compounds are regarded as a much attractive attention for photocatalysts because of

their dispersive and anisotropic band structures, as well as their intrinsic internal electric field. The

driving force of charges separation via internal electric field plays a key role to reduce

recombination rates. In this thesis, it is divided into three parts by using three p-block

photocatalyst based-internal electric field compounds:

Firstly, we have developed a new ferroelectric p-block photocatalyst, Ag10Si4O13, by materials

design and band engineering. Ag10Si4O13 was synthesized by the solid ion-exchange method. It

was found that Ag10Si4O13 has excellent photocatalytic performance towards the degradation of

dyes and phenol over the entire visible light spectrum. The reason behind the efficient activity of

Ag10Si4O13 is attributed to its narrow band gap, dispersive band structure and internal electric field

which has proven theoretically and experimentally.

Secondly, construction unique bismuth silicate based 0D/2D heterogeneous nanostructures was

synthesis by hydrothermal. This heterogeneous structure exhibited much enhanced photocatalytic

activity towards the degradation of Rhodamine B and phenol. This work provides a novel strategy

for adjusting the photoinduced carrier transfer route in the ferroelectric materials and designing

novel photocatalysts with ultrafast charge separation and a large active surface area.

Finally, selective deposition BiOI on the positive polar Bi4Ti3O12 by polarization-adsorption

effect and evaluation of the photocatalytic performance. A complete photodegradation of RhB was

obtained by a specific amount of BiOI interfacing with Bi4Ti3O12 in 12 min under visible light and

47% of phenol has been photodegraded within 1 hour. This is confirmed that selective deposition

on the positive polar surface of ferroelectric produces higher photocatalytic activity rather than

that random heterojunction.

Overall, the results were demonstrated that p-block photocatalyst based-internal electric field

compounds exhibited excellent photocatalytic activity compare with non-p-block photocatalyst

based-internal electric field. The p-block compounds own the unique d10 and sp/p configurations

in its electronic structure and also anisotropic and highly dispersive band structure whereas

internal electric field produced spatially oxidation/reduction (redox) sites on surfaces due to own

high charge separation, also, it can reduce the back-reaction on the surfaces.

Page 4: P-block-based Ferroelectric-photocatalyst compounds

ii

Acknowledgments

I would first like to express my gratitude to my supervisors, Dr. Yi Du and co-supervisor Prof. Shi

Xue Dou, for their support, guidance, encouragement, and advices over four years of my study in

the Institute for Superconducting and Electronic Materials (ISEM) at University of Wollongong. I

am grateful for prof. Weichang Hao Beihang University, Beijing China for his advices. I am also

grateful to staff and technicians for helping me throughout my study, specially, A/Prof. Germanas

Peleckis, Dr Xun Xu and Dr Dongqi Shi and ISEM Administration Officer Mrs Crystal Mahfouz

and AIIM Safety Officer Ms Joanne George and all students in our group for their help especially

long ren. Also, I would like to gratefully acknowledge the financial support from ministry of high

education and scientific research in Iraq to support me throughout my study. Also, I would like to

thank The Electron Microscopy Centre (EMC) and Intelligent Polymer Research Institute (IPRI),

University of Wollongong.

Finally, my thanks are for my parents, my wife, my brothers and my sister for their support and

encourages throughout my study.

Amar Al-keisy

Page 5: P-block-based Ferroelectric-photocatalyst compounds

iii

Certification

I, Amar Hadee Jareeze Al-Keisy, declare that this thesis submitted in fulfilment of the

requirements for the conferral of the degree Doctor philosophy, from the University of Wollongong,

is wholly my own work unless otherwise referenced or acknowledged. This document has not been

submitted for qualifications at any other academic institution.

Amar Hadee Jareeze Al-Keisy

20th May 2018

Page 6: P-block-based Ferroelectric-photocatalyst compounds

iv

List of Abbreviations

Instruments group

AFM : Atomic Force Microscopy

EDS : Energy Dispersive X-ray Spectroscopy

SEM : Scanning Electron Microscope

TEM : Transmission Electron Microscope

XRD : X-ray diffraction

PFM: piezoresponse force microscopy

AFM: Atomic force microscopy

FESEM: The field emission scanning electron microscope

UV-Vis spectroscopy : ultraviolet-visible spectroscopy

Electrochemical group

Ag/AgCl : Silver/ silver chloride reference electrode

NHE: Normal hydrogen electrode

SCE: saturated calomel electrode

Efb: flat-band potentials

Pt: Platinum

calculation group

GGA: generalized gradient approximation

PBE: functional of Perdew, Burke, and Ernzerhof

TMO: transition metal oxide

DFT: Density functional theory

Dye group

CV : Crystal Violet

MB : Methylene Blue

MO: Methyl orange

MB: Methylene blue

RhB: Rhodamin B

Different group

TOC : Total Organic Carbon

AOPs : Advanced Oxidation Processes

CB : Conduction Band

CNTs : Carbon Nanotubes

UV : Ultra Violet

VB : Valance Band

Eg Optical band gap of the semiconductor

FTO: Fluorine-doped tin oxide

TM : Transition metals

Page 7: P-block-based Ferroelectric-photocatalyst compounds

v

Contents

P-block-based Ferroelectric-photocatalyst compounds: Structure, Ferroelectric properties and

photocatalytic performance ..................................................................................................... 1

Abstract .................................................................................................................................... i

Acknowledgments ..................................................................................................................... ii

Certification ............................................................................................................................. iii

List of Abbreviations ................................................................................................................ iv

Contents ................................................................................................................................... v

List of Tables, Figures and Illustrations ................................................................................... viii

Chapter 1 ................................................................................................................................. 1

Introduction .......................................................................................................................... 1

1.1Background and motivation ......................................................................................... 1

1.2 Fundamentals and mechanism of photocatalysis ........................................................ 2

1.3 Mechanism of photodegradation ................................................................................ 4

1.4 Photocatalysis applications ......................................................................................... 6

1.5 Aims and objective ...................................................................................................... 6

1.6 Overview of thesis ...................................................................................................... 7

1.7 References .................................................................................................................. 7

Chapter 2 ................................................................................................................................10

Literature review ..................................................................................................................10

Introduction ....................................................................................................................10

2.1 Challenges and opportunities to enhance and explore photocatalysis ........................10

2.1.1 Band structure engineering and visible light sensitization ........................................10

2.1.2 Enhancement of charge separation and visible light activity in heterojunction

systems. ..........................................................................................................................17

2.1.2.1 Semiconductor-metal composites ................................................................17

2.1.2.2 Semiconductor-Semiconductor composites ..................................................19

2.1.2.3 Carbon Nanostructured- Semiconductors .....................................................22

2.1.2.4 Dye-sensitization Semiconductor..................................................................23

2.1.3 Enhancement of charge separation by internal electric field and facets...................24

2.1.4 Ferroelectric photocatalyst .....................................................................................27

2.2 Summary and opportunities ............................................................................................31

2.3 References ......................................................................................................................31

Chapter 3 ................................................................................................................................39

Experimental techniques ......................................................................................................39

3.1 Materials: ..................................................................................................................39

Page 8: P-block-based Ferroelectric-photocatalyst compounds

vi

3.2 Synthesis....................................................................................................................40

3.2.1 Solid ion-exchange method .................................................................................40

3.2.2 Hydrothermal Method ....................................................................................40

3.2.3 Molten salt synthesis.......................................................................................41

3.2.4 Thin film synthesis ...........................................................................................41

3.3 Characterization: Equipment, instruments, and tools .................................................41

3.3.1 Scanning electron microscopy / Energy dispersive X-ray spectroscopy ................41

3.3.2 Transmission electron microscopy (TEM)/EDS .....................................................42

3.3.3 X-ray powder diffraction .....................................................................................42

3.3.4 Ultraviolet-Visible-Near Infrared (UV-VIS-NIR) Spectrophotometer .....................42

3.3.5 X-ray photoelectron spectroscopy (XPS) ..............................................................43

3.3.6 Atomic force microscopy (AFM) and piezoresponse force microscopy (PFM) ......43

3.3.7 Polarization measurement ..................................................................................44

3.3.8 Photoelectrochemical and electrochemical measurements .................................44

3.3.9 Photoluminescence (PL) spectra..........................................................................45

3.3.10 Surface photovoltage (SPV) ...............................................................................45

3.4 Photocatalytic activity evaluation ...............................................................................45

3.4.1 Determine reaction kinetics kt and quantum efficiency .......................................46

3.5 Programs: analysis .....................................................................................................46

3.6 Theoretical calculations .............................................................................................46

3.7 References ......................................................................................................................47

Chapter 4 ................................................................................................................................49

A ferroelectric photocatalyst Ag10Si4O13 with visible-light photooxidation properties ............49

4.1 Introduction ...............................................................................................................49

4.2 Experimental section .................................................................................................50

4.2.1 Characterizations ................................................................................................50

4.3 Results and discussion ...............................................................................................51

4.4 Conclusions ................................................................................................................68

4.5 References .................................................................................................................69

Chapter 5 ................................................................................................................................71

Enhancement of charge separation in ferroelectric heterogeneous photocatalyst

Bi4(SiO4)3/Bi2SiO5 nanostructures ..........................................................................................71

5.1 Introduction ...............................................................................................................71

5.2 Experimental..............................................................................................................72

5.2.1 Synthesis of bismuth silicate ...............................................................................72

5.2.2 Characterizations ................................................................................................72

5.2.3 Photocatalytic activity evaluation ........................................................................73

Page 9: P-block-based Ferroelectric-photocatalyst compounds

vii

5.2.4 Theoretical calculation ........................................................................................73

5.3 Results and discussion................................................................................................74

5.4 Conclusion .................................................................................................................89

5.5 References .................................................................................................................89

Chapter 6 ................................................................................................................................92

Selective deposition of BiOI on the positive polar surface of Bi4Ti3O12 by the polarization-

adsorption effect and evaluation of the photocatalytic performance ....................................92

6.1 Introduction ...............................................................................................................92

6.2 Experimental..............................................................................................................93

6.2.1 Synthesis of sample.............................................................................................93

6.2.2 Selective photodeposition of metal/metal oxide on BTO plate: ...........................93

6.6.3 Selective deposition of BiOI on BTO plate: ...........................................................94

6.6.4 Characterization..................................................................................................94

6.6.5 Thin film preparation ..........................................................................................94

6.6.6 Photocatalytic activity evaluation ........................................................................95

6.3 Results and Discussion ...............................................................................................96

6.4 Conclusion ...............................................................................................................108

6.5 References ...............................................................................................................108

Chapter 7 ..............................................................................................................................111

Conclusions and future work ..............................................................................................111

7.1 Conclusions ..............................................................................................................111

7.2 Future work .............................................................................................................112

Appendices ...........................................................................................................................113

Appendix 1 .........................................................................................................................113

List of publications .........................................................................................................113

Page 10: P-block-based Ferroelectric-photocatalyst compounds

viii

List of Tables, Figures and Illustrations

Figure 1-1 Schematic diagram of a typical photocatalyst illustrates that the electrons transfer

to the conduction band after absorbing light, leaving a hole in the valence band, and move to

the surface to react with materials have previously been adsorbed on the surface. ................. 3

Figure 1-2 Band-gap positions for photocatalysis in different semiconductors, and their

capability to produce radical species and convert CO2 to hydrocarbon compounds [26]. ......... 4

Figure 1-3 Schematic illustration of degradation process for pollutants (P) by photoinduced

electron and hole in TiO2 [28]. .................................................................................................. 5

Figure 2-1 Schematic illustration of achieving visible light activity in TiO2 by doping with metals

and non-metals [18]. ...............................................................................................................11

Figure 2-2 Density of states (DOS) for different doping elements calculated by FLAPW for non-

metal-doped TiO2: (A) substitutional sites and (B) interstitial sites [6]. ....................................12

Figure 2-3 The DOS of metal-doped TiO2 and pure TiO2. Black solid lines: total DOS (TDOS), and

red solid lines: metal-doped TiO2 with 3d states. The blue dashed line represents the position

of the Fermi level [26]. ............................................................................................................13

Figure 2-4 Band structure (top) and density of states (bottom) of AgMO2 (M = In (right), Ga

(middle), and Al (left)) [31]. .....................................................................................................14

Figure 2-5 Schematic illustration of the electronic structures for BiOX (X = Cl, Br, I) [28, 29]. ...15

Figure 2-6 Band gap values for many photocatalyst compounds [33].......................................16

Figure 2-7 Band gap positions of BiOBr, and Bi24O31Br10 [32]. .................................................16

Figure 2-8 Schematic band diagrams show the mechanisms for the plasmonic effect and the

Schottky junction in composites with n-type and p-type semiconductors [44]. ........................19

Figure 2-9 V Model of semiconductor-semiconductor junction [45].........................................20

Figure 2-10 Photodegradation Rhodamine B (RhB) of over 75% for BiOBr/Bi24O31Br10 is 9.0

times higher than over pure BiOBr or Bi24O31Br10 [49]. .............................................................22

Figure 2-11 MB degradation under visible light of N-doped sites on carbon nanotubes (CNTs)

/TiO2 core/shell nanowire. Inset photograph compares the colour of MB solution (10 ppm)

after 3 h of visible light irradiation in the presence of NCNT/TiO2 nanowires and in the

presence of a P-25 commercial TiO2 photocatalyst [72, 73]. ....................................................23

Figure 2-12 The basic principle of dye-sensitized photocatalytic H2 production from water [77].

...............................................................................................................................................24

Figure 2-13 Crystal structure of Bi2O2[BO2(OH)] showing the internal electric field direction

[84]. ........................................................................................................................................24

Figure 2-14 Schematic illustration of electric field direction and photocurrent for BiOCl (BOC-

001) and BiOCl (BOC-010) samples [87]. ..................................................................................25

Figure 2-15 Photodegradation of bisphenol under visible light by BiOI with (110) and (001)

facets [90]. ..............................................................................................................................26

Figure 2-16 Schematic illustration of the mechanism of ng-CN/BOC [91]. ...............................26

Figure 2-17 (a) IEF in C-doped Bi3O4Cl samples, (b) PL spectra of pure and C-doped Bi3O4Cl, (c)

efficiency of pure Bi3O4Cl, C-doped Bi3O4Cl, TiO2, BiVO4, CdS, C3N4, Fe2O3, and Ag3PO4. d)

Photoredox activity for oxygen and hydrogen evolution on pure and C-doped Bi3O4Cl [92]. ...27

Figure 2-18 Schematic illustrations of the BaTiO3 crystal structure show (A) cubic phase at high

temperature (above the Curie temperature), while (B) and (C) show displacement of the Ti

atom when poling up and down [91]. ......................................................................................28

Page 11: P-block-based Ferroelectric-photocatalyst compounds

ix

Figure 2-19 (a) Schematic illustration of single domain PbTiO3 nanoplate showing the

polarization direction; (b) SEM image of selective photodeposition of Pt and MnOx with

schematic illustration in the inset [112]. ..................................................................................29

Figure 2-20 Schematic diagram (a) showing random movement of photoinduced electron-hole

pair; (b) arrangement of photoinduced electron-hole pairs, with the electrons attracted to the

positive surface of BaTiO3, whereas the holes are attracted to the negative surface of BaTiO3

[119]. ......................................................................................................................................30

Figure 2-21 (a and b) Photodegradation of RhB by BOIO3 samples under visible and UV light,

respectively; (c) Transient photocurrent and (d) photovoltage spectra for V-doped BiOIO3 and

BiOIO3 [120] . ..........................................................................................................................30

Figure 2-22 Schematic diagram of the crystal structure of Ag9(SiO4)2NO3. E is the internal

electric field caused by the polar crystal structure [129]. .........................................................31

Figure 3-1 Photograph of UV-Vix_NIR spectrometer and schematic diagram showing the

geometry in an integrating sphere. .........................................................................................43

Figure 4-1 XRD pattern of the as-prepared Ag10Si4O13. The inset is the crystal structure of

Ag10Si4O13. ...............................................................................................................................52

Figure 4-2 SEM image of the as-prepared Ag10Si4O13 ......................................................53

Figure 4-3 Figure 4-3 (a) TEM image and (b) HETEM image and inset SAED pattern of Ag10Si4O13.

...............................................................................................................................................53

Figure 4-4 Figure 4-3 (a) XPS result of the as-prepared Ag10Si4O13 (b) Ag 3d, (c) O 1s, (d) Si 2p.

The spectra demonstrate that the main peaks correspond to Ag 3d5/2 and Ag 3d3/2, O 1s, and Si

2p orbitals for the Ag10Si4O13. ..................................................................................................54

Figure 4-5 Figure 4-5 EDS spectrum of elements and inset table of elements content

percentage corresponding to the Ag10Si4O13. ...........................................................................55

Figure 4-6 UV-vis diffuse reflectance spectra of Ag10Si4O13. Inset at the top right of (d) is the

derivation of the band gap value; Inset at the bottom left of figure is a photograph of the

photocatalyst. .........................................................................................................................56

Figure 4-7 DFT calculated electronic structure of Ag10Si4O13, in which a dispersive CB and flat VB

can be observed ......................................................................................................................56

Figure 4-8 Partial density of states calculated by DFT. ..................................................57

Figure 4-9 4-9 Mott–Schottky plot of pure Ag10Si4O13. The flat band potential is determined to

be about 0.523 V. The inset is a schematic diagram of the redox potential of Ag10Si4O13,

corresponding to the Mott–Schottky fitting result schematic of band-edge potentials of

Ag10Si4O13. ...............................................................................................................................58

Figure 4-10 UV-Visible absorbance spectra for the photodegradation of (a) MO, (b) RhB and (c)

MB under visible light over Ag10Si4O13 recorded after different degradation times. The insets

shows the color changes of the (a) MO, (b) RhB and (c) MB solutions corresponding to the five

degradation times from 0 min to 40 min .................................................................................59

Figure 4-11 photodegradation rate of organic dyes over Ag10Si4O13 under visible light. ............60

Figure 4-12 Cycling runs in the photodegradation of RhB over Ag10Si4O13. .................61

Figure 4-13 Comparison of the photodegradation percentages of Ag10Si4O13 and N doped TiO2

(N–TiO2) under visible light for different organic dyes. ............................................................61

Figure 4-14 Photodegradation rate of phenol over Ag10Si4O13 under visible light. Data points

represent an average of duplicate measurements, and error bars represent one standard

deviation; the inset shows the kinetic study of the photocatalytic degradation process of

phenol under visible light over Ag10Si4O13. ...............................................................................62

Figure 4-15 High performance liquid chromatography (HPLC) spectra of phenol under visible

light over Ag10Si4O13 ................................................................................................................63

Page 12: P-block-based Ferroelectric-photocatalyst compounds

x

Figure 4-16 Photodegradation of phenol within 20 min over different Ag-incorporated p-block

photocatalysts. References: Ag2CO3[37], Ag3PO4 [38], Ag7Si2O7 [24], Ag10Si4O13 our work. ........64

Figure 4-17 Photocurrent density response with the light on/off over Ag10Si4O13. ....................65

Figure 4-18 Surface photovoltage spectrum of Ag10Si4O13, which shows its highest response in

the visible-light range. .............................................................................................................65

Figure 4-19 P–E hysteresis loop of Ag10Si4O13 at different applied voltages and frequency 1 kHz.

...............................................................................................................................................67

Figure 4-20 Charge density map; the isosurfaces are 0.01e / Å3 of Ag10Si4O13. ..........................67

Figure 4-21 Crystal structure of Ag10Si4O13 to show polarization direction ........................68

Figure 4-22 Three-dimensional chain composed of four SiO4 tetrahedra. Note that the bonds

twist with angles of 166.4° and 143º. ......................................................................................68

Figure 5-1 XRD pattern of as-prepared BSO-HNS ...................................................................74

Figure 5-2 XRD pattern of as-prepared BSO1 .........................................................................75

Figure 5-3 XRD pattern of as-prepared BSO2 ..........................................................................75

Figure 5-4 (a) XPS survey spectrum and (b) EDS of as-prepared BSO-HNS. ...........................76

Figure 5-5 FESEM images of as-prepared BSO-HNS ..............................................................76

Figure 5-6 TEM images of as-prepared BSO-HNS ...................................................................77

Figure 5-7 HRTEM images, inset (1): the SAED pattern, inset (2): HRTEM image side view of as-

prepared BSO-HNS. .................................................................................................................77

Figure 5-8 the top schematic is atomic configuration in the top view of (100) facets and the

bottom schematic is the crystal orientation and polarization direction of as-prepared BSO-HNS.

...............................................................................................................................................78

Figure 5-9 SS-PFM results on the BSO-HNS sample: (a) AFM topography with a white line

cutting cross-section the flake profile with a step height of 15 nm; image scale: 5 × 5 μm2, (b)

PFM amplitude image (white letters a and b represent the flake and substrate area

respectively. ............................................................................................................................79

Figure 5-10 (a) PFM amplitude and (b) phase hysteresis loops for 5 cycles of the BSO-HNS

sample. ...................................................................................................................................80

Figure 5-11 PFM amplitude and (B) phase curves of as-prepared BSO-HNS for 5 cycles of

sweeping triangle/square waves on the substrate, the measurement was also conducted on

the substrate...........................................................................................................................80

Figure 5-12 UV-Vis diffuse reflectance spectra of BSO-HNS, BSO1, and BSO2 with the inset at

the top right showing the derivation of the band-gap values of the as-prepared samples

derived from the diffuse reflectance spectra. ..........................................................................81

Figure 5-13 Adsorption performance of three samples BSO1, BSO-HNS and BSO2 has carried

out in dark in interval 30 min for each RhB and phenol. ..........................................................82

Figure 5-14 UV-Vis absorbance spectra of the photodegradation of RhB dye by BSO-HNS under

UV-Vis light, recorded after different degradation times .........................................................82

Figure 5-15 photodegradation rate of RhB dye by BSO-HNS, BSO1, and BSO2 under UV-Vis

light, with the inset showing a kinetic study of the photocatalytic degradation .......................83

Figure 5-16 photodegradation rate of phenol by BSO-HNS under UV-Vis light, with the inset

showing the UV-Vis absorbance spectra of the degradation of phenol. ...................................84

Figure 5-17 5 recycling runs to test the degradation of RhB by BSO-HNS under UV-Vis light; (b)

XRD patterns before and after 5 cycling runs to test the photodegradation of RhB by BSO-HNS

under UV-Vis light. ..................................................................................................................84

Figure 5-18 XPS spectra of Bi 4f before and after 5 recycling runs in the photodegradation of

RhB by BSO-HNS under UV-Vis light. .......................................................................................85

Figure 5-19 Photocurrent density response under UV-vis light. ................................................86

Page 13: P-block-based Ferroelectric-photocatalyst compounds

xi

Figure 5-20 PL spectra for each BSO2 and BSO-HNS samples. ...............................................86

Figure 5-21 Mott–Schottky plots of BSO-HNS, BSO1, and BSO2. ..........................................87

Figure 5-22 Schematic diagram of the redox potentials of, BSO1, BSO2 and Possible

photocatalytic mechanism of the BSO-HNS sample according to the Mott–Schottky and DRS

spectroscopy results................................................................................................................88

Figure 5-23 Charge density map of (100) facet of BSO-HNS ...................................................88

Figure 6-1 (a) XRD patterns, (b) XPS survey spectra, (c, d, e, and f) high-resolution XPS

spectra for Bi 4f, O 1s, I 3d, and Ti 2p, respectively, for the BTO, BiOI, BTO-BI-1, BTO-BI-2,

and BTO-BI-3 samples. ...........................................................................................................96

Figure 6-2 FESEM images of the (a) BTO,(b) BTO-BI-1, (c) BTO-BI-2, and (d) BTO-BI-3

plates, respectively; (e) a typical single plate of BTO/BiOI; and the EDS mapping of (f) Bi, (g)

I, (h) Ti, and (i) O, and (g) TEM image for a BTO/BiOI heterojunction; (k) the corresponding

SAED pattern, and (l) the corresponding HRTEM image. ........................................................98

Figure 6-3 Photodegradation rates for the RhB dye by all samples under visible light. (b)

Typical visible absorbance spectra of the photodegradation of the RhB dye by BTO-BI-3 under

visible light, recorded after different degradation times. (c) Photodegradation rates of BTO,

BTO-BI-1, and BTO-BI-3 for phenol within 60 min. (d) Adsorption performance of BTO, BTO-

BI-1, BTO-BI-2, and BTO-BI-3 carried out in the dark after an interval of 30 min for RhB and

MO. (e) comparison of photodecolouration rate percentage of the RhB, MB, and MO dyes

achieved by BTO and BTO-BI-3 for different irradiation times under visible light. (f)

Comparison of the photodegradation by BTO-BI-3 and BTO-BI-X for RhB after 10 min. .....100

Figure 6-4 (a-f) SEM image of BTO-BI-X sample, full spectrum EDS mapping of BTO-Bi-X

and EDS mapping of Ti, Bi, I, and O, respectively.................................................................101

Figure 6-5 (a) Scanning electron microscope (SEM) images of BTO plate with photodeposited

Ag on positive potential of polarization. (b) TEM image of deposited Ag on BTO. (c) SAED

pattern of the inside of the yellow rectangle in the TEM image in (b), and (d) SEM image of

photodeposited MnOx on BTO. .............................................................................................102

Figure 6-6 (a) On-Off photocurrent and (b) photovoltage decay for samples under visible light.

(c) UV-vis diffuse reflectance spectra of BTO, BTO-BI-1, BTO-BI-2, and BTO-BI-3 plates,

with the inset at the top right of (c) the derivation of the band gap values of the four samples,

which are derived from the diffuse reflectance spectra. (d) MS plots of BTO and BiOI. .........104

Figure 6-7 (a) On-Off photocurrent and (b) open circuit photovoltage of samples. (c) Nyquist

plots at open-circuit potential under dark and illumination. (d) Mott–Schottky plots for positive

poling, no poling, and negative poling samples. .....................................................................105

Figure 6-8 Schematic diagram of BTO/BiOI thin film and corresponding band structures for

negative poling and positive poling photoanodes. ..................................................................106

Figure 6-9 (a) Bar chart to show role electrons and holes in photodegradation (b) SEM image

and EDS spectrum (inset) of BTO-BI-3 with photoreduction of Ag. ......................................106

Figure 6-10 Mechanism of BTO/BiOI heterojunction: (a) schematic illustration of growth of

BiOI on edge of BTO; (b) possible mechanism of heterojunction; (c) crystal structure for BTO

and BiOI shown at the interface; (d) theoretical calculation of interface band structure. .........107

Table 3-1 all the general chemicals and reagents used in this work. ........................................39

Table 3-2 all the dyes used in this work. ..................................................................................40

Table 3-3 Programs used in this work. .....................................................................................46

Table 4-1 the φ value for different organic pollutants .............................................................63

Page 14: P-block-based Ferroelectric-photocatalyst compounds

1

Chapter 1

Introduction

1.1Background and motivation

Because of the growing world population and increased economic growth, there has been an

increase in water contamination by harmful substances and an increase in carbon dioxide in the air

from burning fossil fuels. Moreover, fossil fuel resources are limited and will be restricted in the

future, but fossil fuels are currently a source of air pollution and also playing a role in global

warming, making the planet less liveable for people in future. The demand for global power is

expected to rise by 30% between today and 2040 [1]. Moreover, by 2050, world population will

much higher. Thus, wastewater will be doubled. Currently, global warming has led to warnings of

disaster if there are no precautions take into consideration. Scientists and engineers around the

world have been seriously concerned with these issues. Therefore, significant efforts have been

made to develop techniques to eliminate these harmful substances, which produced by many

sources, especially industry. It is well known that conventional methods (include adsorption,

coagulation, flocculation, sedimentation, and bio-filtration) cannot degrade recalcitrant

compounds [2]. Thus, some heavy metal ions and toxic organic compounds are still present in

wastewater after treatment, even expensive treatment, and not meet the standards required for the

population [3]. Alternative advanced technologies have been adopted and developed, such as

advanced oxidation processes (AOPs) which have been shown to be promising techniques to break

down the harmful chemicals in wastewater. They can remove and degrade a wide range of organic

components by generating active species radicals. These species attack the organic pollutants in

water and react with them to generate safe materials such as CO2 and H2O. This technique is

considered as expensive for practical application, however, because it uses ultraviolet (UV) light

and ozone or other oxidation species. A few decades ago, among the AOPs, the photocatalysis

process received much attention as a technology to solve the energy crisis and environmental

pollution, because this technique is based on solar energy, which is a permanent and clean source

of energy as well, as it is environmentally safe and renewable.

Photocatalysis is defined as “a photoreaction that is accelerated by the presence of a catalyst.”

Photocatalysts, such as TiO2, have been demonstrated to have significant efficiency in degrading

widespread recalcitrant toxic compounds [4-6]. 120,000 TW (~1000 Wm-2) of the sunlight power

strikes the Earth’s surface, which far exceeds the human demand for power and thus offers free

and clean sustainable energy. The photocatalytic materials are not only applied in

photodegradation and water splitting, but also have antibacterial, self-cleaning, and other

applications. Hydrogen production from water splitting has been considered an attractive, to

environmentally friendly, and sustainable source of energy. Until now, photocatalyst techniques

for practical applications are still under challenge because of many limitations, such as high rates

Page 15: P-block-based Ferroelectric-photocatalyst compounds

2

of recombination of photoinduced carriers, wide band gaps, high cost, and instability. Therefore,

the aims of this project were to develop, enhance, and explore photocatalysts to fulfill the

requirements of practical application.

1.2 Fundamentals and mechanism of photocatalysis

Photocatalysis is a process in materials (including semiconductors, polymers, heterostructures,

etc.) that is able to produce chemical modification or new products on the surface of photocatalyst

materials by absorption of a photon of light. There are two principal steps in photocatalysis: in the

first, the material, normally a semiconductor, is photoexcited to generate electrons and holes, and

in the second, these photoexcited electrons and holes are involved in a series of chemical

reactions. Among these photocatalysts are semiconductor-based photocatalysts. Semiconductor

photocatalysts such as TiO2 were widely studied in the past because of their chemical stability,

superhydrophilicity, strong ability to oxidize organic pollutants, and nontoxicity.

The photocatalysis mechanism starts to work when a photon of light (wavelength, λph = 300-700

nm) is absorbed that has energy (hν) greater than or equal to band-gap energy of a

semiconductor, such as TiO2 [4, 7-9], ZnO [5, 10-12], CdS [13], or WO3 [14-16], and then an

electron in valence band (VB) transit to conduction band (CB) (Equation (1.1)), thus leaving

positive charge in the VB called hole [17, 18]. And then they (photoinduced electrons and holes)

move to the surface of the photocatalyst, the reactions will occur between electrons and holes with

previously adsorbed materials (such as H2O, oxygen, and organic pollutants) to modify them or

produce new chemicals. The electrons in CB (e-CB) act as reductive sites, whereas the holes in the

VB (h+VB) act as oxidative sites.

Photocatalyst hν→ eCB

− + hVB+

In the case where the photogenerated electron-hole recombines on the surface or inside the bulk of

the photocatalyst, the photon energy will be dissipated by emission heat or photons. The photo-

degradation of organic materials occur when powerful reductive/oxidative agents such as holes,

OH•, O2•−, and H2O2 produced in water that generated from reaction holes and electrons with

previously dissolved oxygen and water molecules adsorbed on the surface of photocatalyst

(Equation 1.2). The schematic illustrations in Figure 1-1 shows the process in a typical

photocatalyst [19]. In this case, the conduction band has a negative potential (i.e. reduction

potential over 0 V vs. normal hydrogen electrode( NHE), and the valence band has a positive

potential (i.e. oxidation potential over 1.23 V vs. NHE), so the photocatalyst can convert

(Equations (1.2) and (1.3)) organic compounds and H2O to CO2 and H2O, and H2, respectively

[20-22]. Otherwise, the photocatalyst can only oxidize organic pollutants. Thus, the hydrogen

production or pollutant degradation strongly depends on the band and electronic structure of the

photocatalyst

H2O photocatalyst/ hν→ → O2 + H2

1-2

Organic compounds𝑝ℎ𝑜𝑡𝑜𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡/ ℎ𝜈→ intermediate(s) → CO2 + H2O

1-1

Page 16: P-block-based Ferroelectric-photocatalyst compounds

3

Figure 1-1 Schematic diagram of a typical photocatalyst illustrates that the

electrons transfer to the conduction band after absorbing light, leaving a hole in the

valence band, and move to the surface to react with materials have previously been

adsorbed on the surface.

Generally, the capability to hydrogen evolution or oxygen evolution is strongly depend on

oxidation and reduction potential of VB and CB respectively. Thus, electronic and band structure

are much important to photocatalytic activity. Moreover, there are many factors effect on

photocatalytic activity such as shape and size particles and include internal electric field of

photocatalyst.

The band structure is important to determine the band-gap (Eg) value and the CB and VB

positions of a photocatalyst. The CB and VB arrangement for many photocatalysts is shown in

Figure 1-2. From the Figure, in the case of hydrogen production, it is clear that the electron in the

CB must be higher than the hydrogen reduction potential (0 V on the NHE scale), while the hole

in VB must be below the oxygen oxidization level (1.23 V on the NHE scale) [22]. Materials

design and band engineering can be used to modify the Eg, CB, and VB to absorb visible light in

order to fully harvest solar light and also modify the CB and VB positions to achieve a favourable

level of oxidation potential for the organic pollutants. Highly dispersive CB and VB bands cause

the charge carriers to have much smaller effective masses, which facilitates charge transfer to

surface. Moreover, the natural surface/interface properties of the photocatalyst play an important

role in the photocatalytic activity, because higher surface energy leads to higher photocatalytic

activity [23]. The charge carriers (electrons and holes) should diffuse to the surface of a

photocatalyst in a time shorter than the average lifetime of the electron-hole pair, because

otherwise they will recombine. Therefore, nanostructures offer short distances for the migration of

electron-hole pairs to the surface, and the creation of a high surface area and increasing the density

……..1-3

Page 17: P-block-based Ferroelectric-photocatalyst compounds

4

of reactive sites will also help. Reducing the size of particles cannot always increase their

photocatalytic activity, however. Non-toxic and low-cost photocatalytic materials are also

essential factors, as well as the stability of photocatalyst in a particular practical application. To

overcome all these issues, many efforts were carried out to development, enhancement, and

exploration photocatalysts [24, 25].

Figure 1-2 Band-gap positions for photocatalysis in different semiconductors, and

their capability to produce radical species and convert CO2 to hydrocarbon

compounds [26].

1.3 Mechanism of photodegradation

One major danger facing the environment is pollutants, both organic pollutants and heavy metal

ions. These pollutants are found in water and air, which are waste products coming from various

industries. These pollutants kill marine life and damage the central nervous systems of animals.

Some of these pollutants are highly toxic, such as phenolic compounds, fungicides, herbicides,

dyes. It has been mentioned above that these pollutants cannot be removed by conventional

methods and AOPs can be expensive. Therefore, photocatalysis processing, which is one of AOPs,

is a promising technique to overcome pollution in the environment. It has been demonstrated

completely that the degradation and mineralization pollutants can be achieved under solar light.

The holes in the valence band can directly react with water to produce proton (H+) and the

hydroxy radical OH-. Then, OH- is converted to OH• by reaction with a hole. Also, the holes can

directly react with pollutants in adsorbed water, whereas electrons in the conduction band can

Page 18: P-block-based Ferroelectric-photocatalyst compounds

5

produce superoxide radicals (O2•−) by reacting with dissolved oxygen in water and also produce

H2O2. The holes, OH•, O2•−, and H2O2 are powerful oxidation species toward the degradation

pollutants and other recalcitrant substances. Figure 1-3 shows a schematic illustration of the

degradation process for pollutants (P) by photoinduced electrons and holes in TiO2 [27]. These

species attack organic pollutants in water, which leads to the decomposition and mineralization of

pollutants to simple compounds such as CO2 and H2O. Unfortunately, the efficiency of the

photocatalyst often cannot meet practical requirements because of certain limitations. The major

limitations of a photocatalyst relate to its photocatalytic activity and stability. A high

recombination rate in photoinduced charge carriers leads to high dissipation of light energy and

thus reduces photocatalytic activity. The photoinduced charge carrier lifetime to reach the surface

should be longer than recombination time. Besides wastage from the recombination of

photoinduced charge carriers and instability, the reverse reactions occur on the photocatalyst

surface because of reduction and oxidation sites in the same location on the surface [28], and the

activity of the photocatalyst is also affected by the wettability of the surface.

Figure 1-3 Schematic illustration of degradation process for pollutants (P) by

photoinduced electron and hole in TiO2 [28].

It should be noted that all photocatalysts cannot produce highly oxidative species, in particular,

those with a narrow band gap. The narrow-band-gap semiconductors have high absorption

efficiency toward the visible spectrum of solar light. Nevertheless, the narrow band gaps do not

always match with the redox capabilities of organic pollutants, and a wide band gap cannot absorb

visible solar light, which represents 55% of solar light. For example, the CB must be more

negative than the reduction potential of O2 to form O2•− (O2•− /O2 = -0.28 V vs. NHE at pH = 0),

whereas the VB should be more positive than the oxidation potential of H2O to form OH•

(OH•/H2O= +2.27 V vs. NHE at pH = 0). Generally, visible light active, low charge

recombination, stability, low cost, and suitable band structure for oxidation and reduction potential

Page 19: P-block-based Ferroelectric-photocatalyst compounds

6

cannot all be achieved in a photocatalyst, so it remains a great challenge to apply photocatalysis in

practical applications. Therefore, material design strategies and band structure engineering, as well

as morphology control, have been adopted to meet the requirements of practical applications.

1.4 Photocatalysis applications

Beside pollutant degradation and hydrogen production [34-37], there are many other applications

for photocatalysis, and the most important are listed below. Moreover, it is expected that there will

be more applications in future. It should be noted that the ability to apply photocatalysis in a

particular area mainly depends on the reduction and oxidation capability of the photocatalyst along

with many other factors.

• Purification and disinfection of air and water

Organic pollutants and microorganisms are destroyed when water molecules are decomposed

under the solar light in the presence of a photocatalyst, producing free radicals which can destroy

organic pollutants and microorganisms. The most commonly used photocatalyst in this application

is TiO2, in the form of an ultrathin film of TiO2 coated glass. Outdoor air purification was

achieved by coating buildings with TiO2, and TiO2 under solar light can decompose harmful gases

in water such as H2S and NO [29, 30]. Photoreduction of Cr (VI) by a photocatalyst has been

achieved with complete photoreduction to Cr (III) which has little toxicity [31-34].

• CO2 and N2 fixation

Photocatalysis is a promising technique to produce valuable solar fuels such as CO or CH4, and

NH3 from CO2, water and N2 (Equations (1.4) and (1.5), respectively, because it is simple and uses

solar light as its source of energy [21, 22, 35, 36]. Industrial production of NH3 requires high

temperature and pressure in the presence of a catalyst, but photocatalyst can use solar light to

produce renewable fertilizers under mild conditions. Photoreduction of CO2 is much more

complicated than hydrogen production. Generally, converting CO2 to CH4 requires the

participation of eight electrons and 8 holes, while converting N2 to 2NH3 requires the participation

of six electrons and 6 holes. The band structure is very important for triggering the reaction, and

the CB should be more negative than the reduction potential of N2 or CO2, whereas the VB should

be more positive than the oxidation potential of oxygen evolution.

1

2 N2 +

3

2 H2O

photocatalyst/ hν→ → NH3 +

3

4O2 …(1.4)

CO2 + 2H2O photocatalyst/ hν→ → CH4 + 2O2 …(1.5)

1.5 Aims and objective

The primary objective of this work will be to investigate ferroelectric semiconductors, p-block

compounds, and heterostructures, which are regarded as a new family of photocatalysts. A built-in

internal electric field can serve as an electron-hole separator, thus enhancing photocatalytic

activity. The built-in internal electric field not only achieves charge separation, but also constructs

spatially separated reduction and oxidation sites. The primary aim of this thesis is to study the

crystal and electronic structures of p-block ferroelectric (d10 or d0) semiconductors and study their

photocatalytic activity. The specific aims of the study have been the following:

Page 20: P-block-based Ferroelectric-photocatalyst compounds

7

• Present a literature reviews on developments in photocatalysis and specify the gap in

photocatalysts throughout photocatalytic history.

• Design and apply band engineering to ferroelectric materials consisting of ferroelectric p-block

compounds (d10 or d0 oxide semiconductors).

• Prepare photocatalysts and characterize the materials by different instruments, devices, and

techniques and compare them with computational calculations.

• Evaluate the photocatalytic activity of photocatalysts by the degradation of dye under exposure

to light and study the stability and the mechanism.

• Compare these results with previous results.

Some questions that need to be answered in this thesis:

• Does an internal electric field enhance photocatalytic activity?

• Does nanoscale size have an effect on the polarization value of a ferroelectric?

• Does a ferroelectric/semiconductor interface produce band bending?

• Does a ferroelectric produce different sites for reduction/oxidation?

1.6 Overview of thesis

• Chapter 1 explains the principle and mechanism of photocatalysis

• Chapter 2 reviews reports from the literature that are related to the present research.

• Chapter 3 presents details on experimental methods, materials used in the experiments, materials

preparation, collection and analysis of data , structural characterization methods, electronic

structure characterization methods photocatalytic reactor set-up, photocatalysis evaluation, and

density functional theory (DFT).

• Chapter 4 discusses Ag10Si4O13, a ferroelectric photocatalyst with visible-light photooxidation

properties.

• Chapter 5 discusses the enhancement of charge separation in nanostructures of a ferroelectric

heterogeneous photocatalyst, Bi4(SiO4)3/Bi2SiO5.

• Chapter 6 discusses the preparation and study of Bi4Ti3O12.

• Chapter 7 contains the general summary and conclusions.

1.7 References

1. IEA, Key world energy statistics. 2017.

2. Choi, K.J., S.G. Kim, C.W. Kim, and J.K. Park, Removal efficiencies of endocrine disrupting

chemicals by coagulation/flocculation, ozonation, powdered/granular activated carbon

adsorption, and chlorination. Korean Journal of Chemical Engineering, 2006. 23(3): p. 399-408.

3. UN environmental report 2017. United Nations, 2017.

https://www.unenvironment.org/resources/emissions-gap-report.

4. Asahi, R., T. Morikawa, T. Ohwaki, K. Aoki, and Y. Taga, Visible-Light Photocatalysis in

Nitrogen-Doped Titanium Oxides. Science, 2001. 293(5528): p. 269-271.

5. Ullah, R. and J. Dutta, Photocatalytic degradation of organic dyes with manganese-doped ZnO

nanoparticles. Journal of Hazardous Materials, 2008. 156(1): p. 194-200.

Page 21: P-block-based Ferroelectric-photocatalyst compounds

8

6. Muruganandham, M., R. Amutha, G.-J. Lee, S.-H. Hsieh, J.J. Wu, and M. Sillanpää, Facile

Fabrication of Tunable Bi2O3 Self-Assembly and Its Visible Light Photocatalytic Activity. The

Journal of Physical Chemistry C, 2012. 116(23): p. 12906-12915.

7. Linsebigler, A.L., G. Lu, and J.T. Yates Jr, Photocatalysis on TiO2 surfaces: principles,

mechanisms, and selected results. Chemical reviews, 1995. 95(3): p. 735-758.

8. Schneider, J., M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi, M. Anpo, and D.W.

Bahnemann, Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chemical

Reviews, 2014. 114(19): p. 9919-9986.

9. Etacheri, V., C. Di Valentin, J. Schneider, D. Bahnemann, and S.C. Pillai, Visible-light

activation of TiO2 photocatalysts: Advances in theory and experiments. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 2015. 25: p. 1-29.

10. Hong, D., W. Zang, X. Guo, Y. Fu, H. He, J. Sun, L. Xing, B. Liu, and X. Xue, High Piezo-

photocatalytic Efficiency of CuS/ZnO Nanowires Using Both Solar and Mechanical Energy for

Degrading Organic Dye. ACS Applied Materials & Interfaces, 2016. 8(33): p. 21302-21314.

11. Wang, L., S. Liu, Z. Wang, Y. Zhou, Y. Qin, and Z.L. Wang, Piezotronic Effect Enhanced

Photocatalysis in Strained Anisotropic ZnO/TiO2 Nanoplatelets via Thermal Stress. ACS Nano,

2016. 10(2): p. 2636-2643.

12. Sun, C., Y. Fu, Q. Wang, L. Xing, B. Liu, and X. Xue, Ultrafast piezo-photocatalytic

degradation of organic pollutions by Ag2O/tetrapod-ZnO nanostructures under ultrasonic/UV

exposure. RSC Advances, 2016. 6(90): p. 87446-87453.

13. Zong, X., H. Yan, G. Wu, G. Ma, F. Wen, L. Wang, and C. Li, Enhancement of Photocatalytic H2 Evolution on CdS by Loading MoS2 as Cocatalyst under Visible Light Irradiation. Journal of

the American Chemical Society, 2008. 130(23): p. 7176-7177.

14. Amano, F., E. Ishinaga, and A. Yamakata, Effect of particle size on the photocatalytic activity of

WO3 particles for water oxidation. The Journal of Physical Chemistry C, 2013. 117(44): p.

22584-22590.

15. Aslam, M., I.M.I. Ismail, S. Chandrasekaran, and A. Hameed, Morphology controlled bulk

synthesis of disc-shaped WO3 powder and evaluation of its photocatalytic activity for the

degradation of phenols. J. Hazard. Mater., 2014. 276: p. 120-128.

16. Ashkarran, A., M. Ahadian, and S.M. Ardakani, Synthesis and photocatalytic activity of WO3

nanoparticles prepared by the arc discharge method in deionized water. Nanotechnology, 2008.

19(19): p. 195709. 17. Lee, K.M., C.W. Lai, K.S. Ngai, and J.C. Juan, Recent developments of zinc oxide based

photocatalyst in water treatment technology: A review. Water Research, 2016. 88(Supplement

C): p. 428-448.

18. Daghrir, R., P. Drogui, and D. Robert, Modified TiO2 For Environmental Photocatalytic

Applications: A Review. Industrial & Engineering Chemistry Research, 2013. 52(10): p. 3581-

3599.

19. Hoffmann, M.R., S.T. Martin, W. Choi, and D.W. Bahnemann, Environmental applications of

semiconductor photocatalysis. Chemical reviews, 1995. 95(1): p. 69-96.

20. Chen, X., S. Shen, L. Guo, and S.S. Mao, Semiconductor-based photocatalytic hydrogen

generation. Chemical reviews, 2010. 110(11): p. 6503-6570.

21. Hirakawa, H., M. Hashimoto, Y. Shiraishi, and T. Hirai, Photocatalytic Conversion of Nitrogen

to Ammonia with Water on Surface Oxygen Vacancies of Titanium Dioxide. Journal of the American Chemical Society, 2017. 139(31): p. 10929-10936.

22. Gurudayal, J. Bullock, D.F. Sranko, C.M. Towle, Y. Lum, M. Hettick, M.C. Scott, A. Javey, and

J. Ager, Efficient solar-driven electrochemical CO2 reduction to hydrocarbons and oxygenates.

Energy & Environmental Science, 2017. 10(10): p. 2222-2230.

23. Bai, S., L. Wang, Z. Li, and Y. Xiong, Facet-Engineered Surface and Interface Design of

Photocatalytic Materials. Advanced Science, 2016: p. 1600216-n/a.

24. Kabra, K., R. Chaudhary, and R.L. Sawhney, Treatment of hazardous organic and inorganic

compounds through aqueous-phase photocatalysis: a review. Industrial & engineering chemistry

research, 2004. 43(24): p. 7683-7696.

25. Chong, M.N., B. Jin, C.W. Chow, and C. Saint, Recent developments in photocatalytic water

treatment technology: a review. Water research, 2010. 44(10): p. 2997-3027. 26. Tong, H., S. Ouyang, Y. Bi, N. Umezawa, M. Oshikiri, and J. Ye, Nano-photocatalytic

Materials: Possibilities and Challenges. Advanced Materials, 2012. 24(2): p. 229-251.

27. Mehrjouei, M., S. Müller, and D. Möller, A review on photocatalytic ozonation used for the

treatment of water and wastewater. Chemical Engineering Journal, 2015. 263(Supplement C): p.

209-219.

Page 22: P-block-based Ferroelectric-photocatalyst compounds

9

28. Chong, M.N., B. Jin, C.W.K. Chow, and C. Saint, Recent developments in photocatalytic water

treatment technology: A review. Water Research, 2010. 44(10): p. 2997-3027.

29. Dan, M., Q. Zhang, S. Yu, A. Prakash, Y. Lin, and Y. Zhou, Noble-metal-free MnS/In2S3

composite as highly efficient visible light driven photocatalyst for H2 production from H2S.

Applied Catalysis B: Environmental, 2017. 217: p. 530-539.

30. Patil, S.S., D.R. Patil, S.K. Apte, M.V. Kulkarni, J.D. Ambekar, C.-J. Park, S.W. Gosavi, S.S.

Kolekar, and B.B. Kale, Confinement of Ag3PO4 nanoparticles supported by surface plasmon

resonance of Ag in glass: Efficient nanoscale photocatalyst for solar H2 production from waste

H2S. Applied Catalysis B: Environmental, 2016. 190: p. 75-84.

31. Wang, C.-C., X.-D. Du, J. Li, X.-X. Guo, P. Wang, and J. Zhang, Photocatalytic Cr (VI) reduction in metal-organic frameworks: A mini-review. Applied Catalysis B: Environmental,

2016. 193: p. 198-216.

32. Yu, J., S. Zhuang, X. Xu, W. Zhu, B. Feng, and J. Hu, Photogenerated electron reservoir in

hetero-p–n CuO–ZnO nanocomposite device for visible-light-driven photocatalytic reduction of

aqueous Cr (VI). Journal of Materials Chemistry A, 2015. 3(3): p. 1199-1207.

33. Hu, X., H. Ji, F. Chang, and Y. Luo, Simultaneous photocatalytic Cr (VI) reduction and 2, 4, 6-

TCP oxidation over g-C3N4 under visible light irradiation. Catalysis Today, 2014. 224: p. 34-

40.

34. Wan, Z., G. Zhang, X. Wu, and S. Yin, Novel visible-light-driven Z-scheme Bi12GeO20/g-C3N4

photocatalyst: Oxygen-induced pathway of organic pollutants degradation and proton assisted

electron transfer mechanism of Cr(VI) reduction. Applied Catalysis B: Environmental, 2017. 207: p. 17-26.

35. Tahir, M. and N.S. Amin, Indium-doped TiO2 nanoparticles for photocatalytic CO2 reduction

with H2O vapors to CH4. Applied Catalysis B: Environmental, 2015. 162: p. 98-109.

36. He, Z., J. Tang, J. Shen, J. Chen, and S. Song, Enhancement of photocatalytic reduction of CO2

to CH4 over TiO2 nanosheets by modifying with sulfuric acid. Applied Surface Science, 2016.

364: p. 416-427.

Page 23: P-block-based Ferroelectric-photocatalyst compounds

10

Chapter 2

Literature review

Introduction

The first successful light-to-fuel conversion was carried out by Fujishima and Honda in 1972,

which was based on this photocatalysis technique [1]. This has been considered as the start of

photocatalytic history. The first material used as the photocatalyst was TiO2, even though it is not

a perfect material for photocatalytic applications due to many limitations. One of these limitations

is that TiO2 has a wide band gap at 3-3.2 eV. Unfortunately, only 5% of solar light impinging on

the earth’s surface can be utilized by TiO2 photocatalyst, whereas the visible spectrum forms 40%

of solar light [2-4]. Therefore, great efforts have been made to enhance TiO2 in order to utilize the

visible spectrum by either band engineering to modulate the energy band or other methods. Even

though TiO2 efficiency under visible light has been improved, it is still challenging to apply in

practical applications because the recombination rate of electron-hole pairs is high [5-8].

Moreover, stable semiconductor oxides with visible light activity such as WO3 [9-11] and Bi2O3

[12-14] also have low photocatalytic activity. Semiconductors with visible light activity, in

particular, metal-based sulfide semiconductors such as CdS, have high photocatalytic activity and

are suitable for hydrogen production. Nevertheless, metal sulfides are not stable over the long term

[15-17]. Thus, numerous works have been carried out to find a photocatalyst that fulfils the

requirements of practical applications. Recently, it was observed that there is a significant

enhancement in photocatalytic activity compared with the activity two decades ago. Applications

such as air purifiers and self-cleaning windows that work by photocatalysis are already on the

market, but photodegradation and water splitting are still challenging. This literature review

presents a survey that highlights the achievements and enhancements of photocatalysts.

2.1 Challenges and opportunities to enhance and explore photocatalysis

Although ongoing great efforts have significantly improved photocatalytic activity by providing

large surface area and short pathways for photo-induced charges to reach the surface and also have

modified the band structure, existing photocatalysts still do not meet engineering requirements

such the stability, non-toxicity, visible-light-activity and low cost. These topics will be briefly

covered with a survey on overall strategies for enhancement and exploration of photocatalysis,

which is classified into four parts as follows: 1. band gap engineering and visible light

sensitization, 2. heterojunction systems, 3. photocatalyst built –in internal electric field, and 4.

ferroelectric photocatalysts.

2.1.1 Band structure engineering and visible light sensitization

As part of the enhancement of the photocatalytic activity is increasing visible light activity, in

particular, where photocatalyst based semiconductors have a wide band gap (i.e. use ultraviolet

(UV) light which is only 5% of the light of the sun). The enhancement of visible light activity is

Page 24: P-block-based Ferroelectric-photocatalyst compounds

11

carried out by modifying the band-gap energy of the semiconductor. Thus, semiconductors with a

wide band gap cannot utilize the visible light spectrum. The most stable photocatalysts are

semiconductor oxides such as TiO2, ZnO, Nb2O5, etc., but unfortunately, those photocatalysts have

wide band gaps which are only efficient under UV-light. Element doping of semiconductors is an

effective technique to modify the band gap because it creates impurity levels in the forbidden area

between the conduction band (CB) and the valence band (VB), leading to an upshift in the VB or

downshift in the CB, as shown in Figure 2-1, thus reducing the band-gap energy.

Figure 2-1 Schematic illustration of achieving visible light activity in TiO2 by

doping with metals and non-metals [18].

The most efficient photocatalysis reported is for TiO2 (band-gap energy, Eg = 3.2 eV), which is

highly efficient towards water splitting and organic pollutant degradation. TiO2 doped with

Nitrogen atoms is much more effective with respect to visible light activity because the atomic

size is comparable to that of oxygen, and it has low ionization energy and high stability [19]. On

the other hand, doping with fluorine does not decrease the band gap. The non-metal element

dopants (carbon, phosphorus and sulfur) have shown positive results for visible light activity.

Doping with non-metal elements, not only increases visible light activity, but also creates trapping

sites within the VB and CB that increase the lifetime of photoinduced charge carriers. It should be

noted that the requirements on the doping are very strict with respect to atomic size, and the non-

metals elements rarely have an affect on the CB [20]. The density of states calculations (as shown

in Figure 2-2) reveal that the substitutions of non-metal elements (N, P, S, C) for O can lead to the

mixing of non-metal 2p orbitals with O 2p and induce upward shifting of the VB. Co-doping with

non-metals such as N and F in TiO2 leads to higher photocatalytic activity than for N- or F-doping

separately, because N-doping narrows the band gap, whereas F promotes charges separation.

Moreover, the metal element dopants such as V, Fe, Cr, and Co have been shown to increase the

visible light activity of TiO2 as well as enhancing the photocatalytic activity [21]. Theoretical

calculations have shown that metal-doping in TiO2 such as with Sn4+ can create impurity levels in

the lower CB of TiO2 and are considered as electron acceptor levels, whereas V+4 is considered as

an electron donor level in the upper VB of TiO2, as shown in Figure 2-3.

Page 25: P-block-based Ferroelectric-photocatalyst compounds

12

Figure 2-2 Density of states (DOS) for different doping elements calculated by

FLAPW for non-metal-doped TiO2: (A) substitutional sites and (B) interstitial sites

[6].

In addition, the impurity energy levels in the lower CB or upper VB can not only serve to make

the semiconductor visible-light-active, but also serve as electron and hole traps. Therefore, metal-

doping of TiO2 can enhance the photocatalytic activity under visible light. It should be noted that

not all metal doping can enhance the photocatalytic activity because metals can sometimes serve

as recombination sites as in the case of Al. In such a way, the other photocatalysis capable

semiconductor oxides with wide band gaps (such as ZnO) can also benefit from the same strategy

as doped TiO2, but there are some differences in the electronic structure. Manganese-doped ZnO

has exhibited enhancement of optical absorption under visible light, and the results show higher

photocatalytic activity for photodegrading methylene blue (MB) than undoped-ZnO [22]. In

general, element doped semiconductors exhibit a change in photocatalytic activity due to changes

in the energy band structure [23]. However, simple metal oxides or binary oxides have energy

band structures that are not easy to modify because they can only accommodate limited dopant

content. Multi-metal oxides such as Ag+ and Bi3+ based compounds have recently become more

capable of visible light activity by completely or partially replacing one element by another

element.

The electronic configurations containing the cations with d10 or d10s2 orbitals can introduce

occupied d- or s-states into the VB, which can hybridize with O 2p states to produce up-shift of

the valence band maximum (VBM). Whereas, the s, p states from metal (M) contribute to down-

shifting of the energy level of the conduction band minimum (CBM). For example, in the case of

Ag-containing multi-metal oxides, AgMO2 (M = Al, Ga, In), the band structures and the DOS

have been calculated by density functional theory (DFT). The CBM is mainly constructed from

the Ag 5s5p and M s,p states, whereas all the VBs consist of the Ag 4d and O 2p states.

Hybridization occurs between the Ag 5s5p and M s,p states at the CBM, which enhances visible

light activity. Thus, the band gaps of AgAlO2, AgGaO2, and AgInO2 are changed to 2.95, 2.18,

and 1.90 eV, respectively [24]. Similarly, the electronic structure of Ag2MO4 (M = Cr, Mo, W) is

shown in Figure 2-4, where visible light activity is exhibited by changing the metal from Mo → W

Page 26: P-block-based Ferroelectric-photocatalyst compounds

13

→ Cr, where the band gaps are 3.4, 3.1 and 1.8 eV for Ag2MoO4, Ag2WO4, and Ag2CrO4,

respectively [25].

Figure 2-3 The DOS of metal-doped TiO2 and pure TiO2. Black solid lines: total DOS (TDOS), and red solid lines: metal-doped TiO2 with 3d states. The blue

dashed line represents the position of the Fermi level [26].

Page 27: P-block-based Ferroelectric-photocatalyst compounds

14

Figure 2-4 Band structure (top) and density of states (bottom) of AgMO2 (M = In

(right), Ga (middle), and Al (left)) [31].

The Bi3+ based compounds, in particularly, bismuth oxyhalide (BiOX, X = Cl, Br, I) have

demonstrated excellent photocatalytic activity towards pollutant degradation. BiOCl possesses a

wide band gap, however, and thus, the band gap has been reduced by partially changing the

halogen atoms from Cl → Br → I. In general, the band gap decreases with the electronegativity of

halogen elements in compounds (Figure 2-5). Therefore, visible light activity has been achieved,

resulting in enhanced photocatalytic activity. For example, BiOClxBr1–x (0 ≤ x ≤ 1) solid solutions

have successfully fabricated by the hydrothermal method, and the band gap has been reduced

from 3.39 to 2.78 eV by decreasing the ratio of Cl/Br. aThe BiOCl0.5Br0.5 exhibited higher

photodegradation efficiency towards rhodamine B (RhB) [27].

Page 28: P-block-based Ferroelectric-photocatalyst compounds

15

Figure 2-5 Schematic illustration of the electronic structures for BiOX (X = Cl, Br,

I) [28, 29].

The p-block cations with a d10 electronic structure such as Bi2InNbO7 have been given enhanced

photocatalytic activity by completely substituting Ga or Al for In, and the band gaps have been

estimated: Bi2InNbO7 (3.01 eV), Bi2AlNbO7 (2.79 eV), Bi2GaNbO7 (2.67 eV) [30]. In general, it

is easier to modify the band structures of ternary oxides than binary oxides by doping or complete

substitution with various metals. Figure 2-6 shows band-gap values for many photocatalysis

compounds. In fact, the p-block compounds with a d0 or d10 electronic structure have been

significantly more attractive for research because they exhibit excellent photocatalytic activity.

These compounds possess highly dispersive band structures, since the bands consist of anisotropic

p and sp states and the d orbitals, either empty or filled, do not not participate in the band

structure. The photogenerated charges thus possess high mobility, compared to partially filled d

electronic configurations that have flat conduction bands. Hence, the photocatalytic activity has

been significantly enhanced with visible-light-active and unique band structures. New families

with the configuration BixOyXz (X = Cl, Br, I, x, y, and z =1, 2, 3,… 31) were explored as

promising photocatalysts, such as Bi24O31Cl10 [31] and Bi24O31Br10 [32]. The CB of Bi24O31Br10

has been measured and calculated by Mott-Schoktty experiments and DFT, respectively.

Page 29: P-block-based Ferroelectric-photocatalyst compounds

16

Figure 2-6 Band gap values for many photocatalyst compounds [33]

It has been proven that, when the CB is upshifted to a negative value vs. normal hydrogen

electrode (NHE), Bi24O31Br10 has the capability to photocatalyze H2 evolution from water under

visible light, as shown in Figure 2-7. Modifying BixOyXz does not only modulate the band gap, but

also can produce a band-gap upshift [34-36]. Bi5O7Br has been demonstrated excellent activity

than BiOBr to photodegradation of dye molecules. This is because the conduction band became

more negatively with increasing Bi content and then the conduction possess ability to reduce

oxygen molecular to produce •O2. The Figure 2-8 show that Bi24O31Cl10 has much producing of

•O2 compare with BiOCl whereas Bi24O31Br10 has shown capability to produce H2 for water

splitting [37].

Nevertheless, the recombination rate is still high, and there is too much chemical instability to

apply it in practical application. Great efforts have been made to increase charge separation and

visible light activity by combining two materials, which will be discussed in the following ections.

Figure 2-7 Band gap positions of BiOBr, and Bi24O31Br10 [32].

Page 30: P-block-based Ferroelectric-photocatalyst compounds

17

Figure 2-8 Oxygen activation activity for O2•−and •OH generation by Bi24O31Cl10 and BiOCl under

visible light irradiation. shown high O2•− production by Bi24O31Cl10 compare with BiOCl [35]

2.1.2 Enhancement of charge separation and visible light activity in

heterojunction systems.

A semiconductor in contact with with other semiconductors, or molecules, dyes, and metals, can

show enhanced photocatalytic activity compared with each component separately. This system is

called a heterojunction structure, which can promote charge separation and better utilization of the

solar spectrum. The band bending generated when a semiconductor is in contact with some other

material causes bending in the energy levels of the VB and CB for the semiconductor. The

bending results from the different Fermi levels, and thus, in the contact region, the electrons will

redistribute themselves between the two materials until the Fermi levels are aligned at equilibrium.

The energy levels for the semiconductor will bend upward or downward depending on the

differences in the Fermi level. There are many systems of heterojunction according to the

materials combined with semiconductors, which are classified in the following sections.

2.1.2.1 Semiconductor-metal composites

One of the actions to promote photocatalysis is coupling a semiconductor (S) to a metal (M),

which results in visible light activity and electron-hole separation. Several reports have shown that

noble metals on the nanoscale (e.g. Ag, Pt, Pd, and Au) coupled with a photocatalyst (e.g. TiO2)

can enhance the photocatalytic activity.

Composites such as Au/TiO2 Ag/TiO2, Au/ZnO, and Ag/ZnO have shown high photocatalytic

performance under visible light despite the wide band gap of the semiconductor [38]. Basically,

the phenomenon of localized surface plasmon resonance (LSPR) occurs when metal nanoparticles

(NPs) (especially Au and Ag) are illuminated by light with a frequency of that matches the natural

frequency of the surface electrons, which oscillate to resist the restoring force of positive nuclei,

but coherently oscillate in phase with the electromagnetic field of the incident light [39]. The

photoinduced hot electrons can be injected from Au into the CB for an n-type semiconductor and

hot holes can be injected into the VB for a p-type semiconductor, which thus can promote

electron-hole separation and then enhance the visible light activity in the composite. Generally,

Page 31: P-block-based Ferroelectric-photocatalyst compounds

18

the noble metals are mostly used to support a wide range of metal oxides and ternary metal oxides.

For example, Ag/BiOCl/Pd plasmonic photocatalysts have shown enhanced photocatalytic

performance. The photoinduced hot holes in the Ag nanoparticles under visible light can be

injected into the VB of BiOCl. In contrast, Pd works by hole trapping that is photoinduced by

BiOCl under UV light. Thus, this material can utilize full spectrum with high charge separation.

Although BiOCl possesses a wide band gap, it can work under visible light via plasmonic

resonance as well as the reduction/oxidation potentials determined by the positions of the VB and

CB of BiOCl. Therefore, there are two phenomena that can support enhancement: the

metal/photocatalyst plasmonic effect and the Schottky junction. The plasmonic effect can

photogenerate hot charge carriers and inject them into the photocatalyst, whereas the Schottky

junction can work as a sink to accept charge carriers from the photocatalyst [40]. Figure 2-8 shows

Schematic band diagrams for the mechanisms of the plasmonic effect and the Schottky junction in

a composite of an n-type and a p-type semiconductor. Another bismuth oxyhalide, BiOX (X = Cl,

Br, I), was also loaded with noble metals (Rh, Pt and Pd) It followed the order Pt > Pd > Rh for

UV and Rh > Pt > Pd for visible light irradiation.[41]. The results showed significantly enhanced

activity. Similarly, silver-based compounds, Ag@AgBr [42], Ag/Ag3PO4 [43], and M/Ag3PO4 (M

= Pt, Pd, Au) [44], also showed increased activity. Indeed, the LSPR is significantly affected by

the particle size and shape, and the carrier concentration of nanoparticles. Metal-photocatalyst

activity is also affected by many factors such as the work function of the metal Schottky barrier at

the interface and the electron affinity of the semiconductor [40].

Page 32: P-block-based Ferroelectric-photocatalyst compounds

19

Figure 2-8 Schematic band diagrams show the mechanisms for the plasmonic

effect and the Schottky junction in composites with n-type and p-type

semiconductors [45].

2.1.2.2 Semiconductor-Semiconductor composites

Coupling a semiconductor (S) with another semiconductor (S) can enhance charge separation and

also improve the visible light activity if one of them absorbs visible light. As can be seen Figure 2-

9, high charge separation is expected because the electrons and holes will be driven in different

direction. The electron in the CB of S(A) will be injected into the CB of S(B) which has a lower

negative potential, and the hole will be injected from the VB of S(B) into the VB of S(A) which

has a lower positive potential. This type of heterojunction is called type I, where two

semiconductorsare of the same type, for example, n-n type or p-p type.

Page 33: P-block-based Ferroelectric-photocatalyst compounds

20

Figure 2-9 V Model of semiconductor-semiconductor junction [46].

Thus, the electrons and holes will be separated. In order to absorb the full light spectrum, a narrow

band gap S(A) (such as CdS and C3N4) is combined with a wide band gap S(B) (such as TiO2 and

ZnO). Thus, the wide spectrum can be absorbed, and there is higher charge separation compared

with each photocatalyst separately. Numerous efforts were previously carried out to enhance the

photocatalytic activity of the TiO2-based heterojunction, C3N4/TiO2 [47]. Such heterojunctions

have shown higher photocatalytic activity than TiO2, because C3N4 absorbs visible light and has a

CB higher than that of TiO2 but possesses a high recombination rate. Therefore, the electrons from

C3N4 which are photoexcited from visible light can transfer and migrate to the CB of TiO2, giving

this heterojunction enhanced photocatalytic activity. There are many reports on this type of

heterojunction based TiO2 such as SnO2/TiO2 [48], or CdS/TiO2 [49]. It should be noticed that the

S-S composites must have higher interfacing (overlapping) between S-S to enhance charge

separation, especially where the difference in lattice spacing is small between two semiconductors.

Therefore, electrons and holes can be injected into another semiconductor rather than

recombining. According to the previously reported results that heterostructured photocatalysts

containing the same elements should be preferred. For example, heterojunctions with the same

elements, such as BiOBr/Bi24O31Br10 (Figure 2-10) and Bi3.64Mo0.36O6.55/Bi2MoO6 [50, 51], or two

of the same elements, such as Bi2O3–Bi2WO6, BiOI/BiOBr, and BiIO4/Bi2WO6 [52-54], or one

same element [55], such as Bi2S3/BiOBr have demonstrated higher photocatalytic activity

compared with a heterojunction containing materials with high mismatch in the band structure and

lattice. This is because poor interfacing between two semiconductors leads to ineffective electron

transfer pathways. The semiconductor p–n junction (type II) has high charge carrier separation

compared with type I. In this type, the space charge region (depletion layer) at the interface

between the two types of semiconductors exists because the electronsm which are majority

carriers in n-type semiconductor, diffuse inside the p-type semiconductor leaving positive charges,

whereas, on the other side, the holes, which are majority carriers in p-type semiconductor, diffuse

inside the n-type semiconductor leaving negative charges. The diffusion continues until the Fermi

level is aligned for two sides. Thus, the p-n junction creates band bending and an internal electric

field, which are due to powerful charge separation. The photoinduced electrons under light in the

Page 34: P-block-based Ferroelectric-photocatalyst compounds

21

space charge region will be driven under an electric field to the positive region of the n-type side.

Meanwhile, the photoinduced holes will be driven to the negative region of the p-type side. The

photoinduced electrons migrate to the CB of the n-type side, whereas the holes migrate to the VB

of the p-type side, and thus this type of heterojunction offers longer charge carrier lifetimes,

spatial charge carrier separation, and a wide absorption spectrum. As a result, this type of

heterojunction is perfect for practical application. Recently, nanotechnology has offered new

strategies to enhance photocatalytic activity. Two-dimensional (2D) nanosheets have been

demonstrated to have a large surface-to-volume ratio and can provide good platforms on their

surfaces to grow materials with different dimensions (such as 0, 1, and 2D). One-dimensional (1D)

nanostructures such as wires, tubes, and belts have recently received as considerable attention

because they are attractive for photocatalytic application. Compared with other dimensions, the 1D

types feature excellent charge transport and large surface area. 1D TiO2 has been shown to have

efficient electron-hole separation because of superior charge transport. Zero-dimensional (0D)

CdS sensitized 1D TiO2 nanotube arrays have synthesised and exhibited high photocatalytic

activity [56, 57]. SnS2/ g-C3N4 nanosheets as 2D/2D heterojunction photocatalysts have

demonstrated significant enhancement toward Methylene blue (MO) under visible light, and the

photodegradation percentage increased from ∼45.5% and ∼35.4% for C3N4 and SnS2,

respectively, to ∼95.0% for SnS2/ g-C3N4. In this form, the larger contact area exhibits better

charge separation compared with point-to-area as in 0D/2D[58]. Recently, constructing ternary

composites can produce multiple pathways and channels to transfer electrons and holes, and thus

achieve separation. For example, Au nanoparticle (NP) decorated 2D/2D Bi2WO6/TiO2

heterojunctions have demonstrated excellent enhancement of photocatalytic activity. Nevertheless,

ternary composites have been considered less interesting because of their expensive preparation

and lower stability [59]. Ag-based oxides have high photocatalytic activity. Generally, most Ag-

based oxides are responsive in the visible light spectrum, but the main reason why these

photocatalysts lose their activity is because they easily undergo photocorrosion (Ag++ e−→ Ag).

Recently, reports have shown that Ag-based heterojunctions can decrease the photocorrosion. The

photocatalytic activity of AgBr/Ag2CO3 heterojunctions was demonstrated to be higher compared

with Ag2CO3 alone and was maintained after 4 recycling runs [60], while Ag2O/Ag2CO3

heterostructures have shown very much higher and more stable photocatalytic activity than

Ag2CO3 and Ag2O, respectively [61]. Interestingly, the built-in electrical potential found in p-n

heterojunctions has more advantages than in p-p and n-n heterojunctions because there is more

effective charge separation and migration in the presence of the local electric field in the region

between the two different types of semiconductors (p-type and n-type). Ag2O/Bi2O2CO3 p-n

heterojunction [62] photocatalysts exhibited higher photocatalytic activity and greater stability

than pure Bi2O2CO3, Bi2S3/Bi2O2CO3 n-n heterojunctions [63], and Ag2O/Ag2CO3 p-p

heterojunctions [64]. In a p-type semiconductor, the transfer of electrons from n-type to p-type and

holes from p-type to n-type can generate a space charge region, and therefore, a local electric field

is formed between the two types of semiconductors. This electric field can assist in charge

separation and thus higher photocatalytic activity [65].

Page 35: P-block-based Ferroelectric-photocatalyst compounds

22

Figure 2-10 Photodegradation Rhodamine B (RhB) of over 75% for

BiOBr/Bi24O31Br10 is 9.0 times higher than over pure BiOBr or Bi24O31Br10 [49].

2.1.2.3 Carbon Nanostructured- Semiconductors

It has been shown in previous reports that carbon nanotubes (CNTs) not only have useful

mechanical properties and a large specific surface area, but also unique electronic and optical

structures. TiO2-CNT heterojunctions have demonstrated excellent charge separation because the

CNT has a large electron-storage capacity to accept electrons from the semiconductor, thus

suppressing the recombination rate, as shown in Figure 2-11. In addition, CNT has a large surface

area, which provides adsorption capacity toward specific species.

The one-dimensional (1-D) morphology of CNTs as a matrix core/shell architecture leads to

enhanced facilitation of electronic transport and results in improved quantum efficiency in

photocatalysis. On the other hand, 2D graphene (G), graphene oxide (GO), and reduced graphene

oxide (RGO or rGO) have shown excellent physical and chemical properties and are promising

materials for photocatalytic activity compared with carbon forms such as 0D fullerenes, 1D

nanotubes, and 3D graphite. There is an enormous literature on graphene–semiconductor

nanocomposites, and they all have enhanced photocatalytic activity. On the other hand, most

silver-based photocatalysts are sensitive to visible light and have high photocatalytic activity, but

they are subject to high instability. Thus, much research has been done to suppress the

decomposition of silver components by coupling them with different forms of carbon.

CNT/Ag3PO4, GO/Ag/AgX (X = Cl, Br), and GO–Ag3PO4 heterojunctions demonstrated higher

quantum efficiency toward dye degradation and better stability compared with only silver

components [66-68]. Ag/Ag-based photocatalysts could also show some other limitations such as

low adherence, while hybrid Ag3PO4/RGO/Ag heterostructured photocatalyst showed excellent

Page 36: P-block-based Ferroelectric-photocatalyst compounds

23

quantum efficiency and better stability than Ag3PO4/RGO or Ag/Ag3PO4 [69-71]. On the other

hand, in the case of Bi-based photocatalysis, rGO loaded BiOI photocatalyst exhibited enhanced

efficiency with increased loading of rGO [72], and hence, 2D carbon can play an important role in

transporting electrons far away from the CB of the photocatalyst and support, thus enhancing

charge separation.

Figure 2-11 MB degradation under visible light of N-

doped sites on carbon nanotubes (CNTs) /TiO2

core/shell nanowire. Inset photograph compares the

colour of MB solution (10 ppm) after 3 h of visible

light irradiation in the presence of NCNT/TiO2

nanowires and in the presence of a P-25 commercial

TiO2 photocatalyst [73, 74].

0D/2D CdS/Graphene (G) composites have shown excellent quantum efficiency in hydrogen

production under visible light [75]. The G serves as efficient electron collector and transporter to

reduction sites because of its excellent electron conductivity and unique 2D structure, and it also

suppresses the photocorrosion of CdS. Amorphous silver silicates/ultrathin 0D/2D g-C3N4

heterojunction composites have been fabricated, and the enhancement was much higher than for g-

C3N4 alone [76], whereas 0D/0D-2D CdS–Pt–G nanocomposites significantly enhanced the

production of hydrogen under visible light [77].

2.1.3 Dye-sensitization Semiconductor

Dye sensitization, more interestingly, can extend light absorption by utilizing longer wavelengths

in the spectrum, thus enhancing the quantum efficiency of a photocatalyst, and in particular,

supporting a wide band-gap photocatalyst for either hydrogen production or photodegradation.

Basically, previously adsorbed dye on photocatalyst surface results in dye cationic radicals that

can be excited by a photon. The dye generates electrons in the lowest unoccupied molecular

orbital (LUMO), which can be trapped by the conduction band of the photocatalyst (Figure 2-12)

and then react with adsorbed O2 and H2O on the photocatalyst surface to produce radical species

or water splitting. For example, because Bi24O31Cl10 and rhodamine B (RhB) are compatible in

energy level, the Bi24O31Cl10 shows excellent photocatalytic activity under visible light towards

Page 37: P-block-based Ferroelectric-photocatalyst compounds

24

the degradation of RhB, which is promoted by dye sensitization [31]. The same is true of the

RhB–BiOCl system [15] and also rhodamine B with K4Nb6O17 [78] and Nb2O5 [79].

Figure 2-12 The basic principle of dye-sensitized photocatalytic H2 production

from water [78].

Generally, dye-sensitization photocatalysts offers the choice of a variety of dyes compared with

metal nanoparticles that are loaded on a photocatalyst, and each dye possesses different properties

that could be suitable for a photocatalyst. The dye should have a wide absorption spectrum, a

higher energy potential (LUMO) than the CB of the photocatalyst, and high affinity between dye

and photocatalyst [80-82]. Dye-sensitization photocatalysts are utilized for photodegradation of a

wide range of pollutants such as benzyl alcohol, chlorophenol, and phenol [83].

2.1.4 Enhancement of charge separation by internal electric field and facets

Recently, bismuth-based layered structures with internal electric field have demonstrated high

quantum efficiency for photodegradation of organic pollutants such as bismuth oxyhalides. The

visible light activity in these components results from hybridizing O 2p and Bi 6s orbitals in the

VB, whereas the empty Bi 6p orbital is the main constituent of the CB [84].

Figure 2-13 Crystal structure of Bi2O2[BO2(OH)] showing the internal electric field

direction [85].

Page 38: P-block-based Ferroelectric-photocatalyst compounds

25

The crystal structures of these components are constructed from anisotropic [Bi2O2]2+ layers

intercalated by X− ions (X = F, Cl, Br, I), which generates an internal electric field (IEF) between

the layers and thus can enhance charge separation [37, 86]. This IEF is generated between [Bi2O2]

slabs, which have strong covalent bonding with the halogen double slabs which act as a weak

interlayer (only van der Waals forces) along the c-axis.

Spatial charge separation generates different sites for reduction/oxidation and thus suppresses

back-reaction of intermediates and electron-hole recombination. The higher photocatalytic activity

of Bi2O2[BO2(OH)] than Bi2O2CO3 is because Bi2O2[BO2(OH)] (Figure 2-13) has an internal

electric field (IEF), and its sheet structure has shown higher efficiency than irregular structures

even though they have a higher specific surface area [85].

Figure 2-14 Schematic illustration of electric field direction and photocurrent for

BiOCl (BOC-001) and BiOCl (BOC-010) samples [87].

Recently, it has been proved that facets affect the photocatalytic activity, depending on the

photocatalyst surface. It has been demonstrated that the surface atomic configuration due to

different facets leads different photocatalytic activities. By controlling the exposed facets of

BiOCl single crystals with different IEF directions, the different photocatalytic activity of the

(001) and (010) planes could be demonstrated. The photocatalytic activity of (001) facets

benefited from higher charge separation, as illustrated in Figures 2-14 and 2-15. The IEF long

(001) facet favours charge separation and short diffusion distances of charge carriers to the surface

[87]. In another example, anatase TiO2 with {001} facets exhibited a more highly reactive surface

than with (101) facet because TiO2 with (001) facets has higher surface energy [88]. This is also

true of Ag-based photocatalysts, for example, Ag3PO4 with {011} facets exhibited higher quantum

efficiency compared with (100) facets due to their higher surface energy [89]. Similarly, in Bi-

based photocatalysts, BiOI with (110) facets exhibited much higher photocatalytic activity than

with {001} facets towards the degradation of bisphenol under visible light [90]. Facet engineering

helps researchers to understand the photocatalysis mechanism and follow the charge carrier

pathways. In addition, a heterojunction structure with certain facets exposed makes it much easier

to study the mechanism of photocatalysis. A certain facet interface in heterojunctions can serve to

promote higher charge separation compared with a disordered interface. In addition, the particular

facet interface can make it easy study the mechanism involved, because otherwise the

heterojunctions exhibit much complexity in their mechanisms. Recently, other facet

heterojunctions have been widely reported and promoted the study photocatalytic behaviour at

Page 39: P-block-based Ferroelectric-photocatalyst compounds

26

interfaces with different facets. Moreover, electric fields built into composites have demonstrated

different activity. For example, g-C3N4/BiOCl (ng-CN/BOC) heterostructures have exhibited

different photocatalytic activity depending on the IEF direction.

Figure 2-15 Photodegradation of bisphenol under visible light by BiOI

with (110) and (001) facets [90].

Figure 2-16 Schematic illustration of the mechanism of ng-CN/BOC [91].

The separation of electrons and holes is more efficient with an in-plane internal electric than an

out-of-plans. CN/BOC-010 composites demonstrate more highly efficient activity than CN/BOC-

001 composites (Figure 2-16). The IEF of bismuth oxyhalide is not sufficient to separate charges,

however, although theoretical calculations suggested that the IEF of C-doped Bi3O4Cl could be

magnified by many times. It was experimentally proven that C-doped Bi3O4Cl showed an

increased IEF 126 times greater than that of pure Bi3O4Cl (Figure 2-17), and the enhancement in

charge separation efficiency was 80% [92]. Also, the adsorption can demonstrate different

activity. Exposed facets with a high percentage of oxygen termination show high activity towards

H+ under acidic conditions. BiOCl with (001) facets showed higher adsorption activity compared

with the (110) surface due to the high percentage of oxygen termination.

Page 40: P-block-based Ferroelectric-photocatalyst compounds

27

Figure 2-17 (a) IEF in C-doped Bi3O4Cl samples, (b) PL spectra of pure and C-

doped Bi3O4Cl, (c) efficiency of pure Bi3O4Cl, C-doped Bi3O4Cl, TiO2, BiVO4,

CdS, C3N4, Fe2O3, and Ag3PO4. d) Photoredox activity for oxygen and hydrogen

evolution on pure and C-doped Bi3O4Cl [92].

Employing crystal facet engineering to design hybrid single crystal semiconductor-semiconductor

sheets is an important way to enhance photocatalytic activity. It has been confirmed that

BiOI(001)/BiOCl(010) has higher activity than BiOI(001)/BiOCl(001) due to low mismatch of the

lattice parameters, thus facilitating electron transfer pathway[93]. The facet adjustment of each

semiconductors in heterojunction could produce more benefit for maximize the photocatalytic

performance. Although this system is very complexity to synthesis, some facet-controlled

heterojunctions were prepared, the interfacing with certain facet for each photocatalyst is not main

difficultly but also band structure. Band alignment at interfacing can influence on charge transfer,

for example, in Cu2O, the face {001} is much higher work function than {111}. Therefore, the

electrons and holes will be accumulated at {001} and {111} respectively. Deposited Pd on Cu2O

{111} is not useful because Cu2O {111} work function is lower than Pd whereas Cu2O {001} is

higher than Pd, therefore, Pd can be used as hole trapping for Cu2O {001} [94].

2.1.5 Ferroelectric photocatalyst

Photocatalysis based on built-in spontaneous electric polarization that is provided by

ferroelectric materials was proposed to promote spatial charge separation, which provides a

driving force to move electrons and holes in opposite directions, forming two types of sites, one

for reduction and the other for oxidation. Therefore, a photocatalyst with built in spontaneous

electric polarization features suppressed charge recombination and drawback reaction. Also,

similar behaviour to ferroelectric occurs in pyroelectric, piezoelectric, semiconducting, and

electrooptic materials.

The ferroelectric materials have spontaneous polarization arising from the non-centro-symmetric

nature of the crystal structure, meaning that the crystal structure is non-zero neutral because the

positive and negative charges have non-symmetric positions. This polarization is stable and

permanent over a wide range of cruel thermal and chemical environments. The polarization

Page 41: P-block-based Ferroelectric-photocatalyst compounds

28

direction is reversible by an external electric field, causing rearrangement of ions in the crystal

structure. Figure 2-18 shows the crystal structure of BaTiO3 under different conditions[91].

Figure 2-18 Schematic illustrations of the BaTiO3 crystal structure show (A)

cubic phase at high temperature (above the Curie temperature), while (B) and

(C) show displacement of the Ti atom when poling up and down [91].

In ferroelectric materials, some regions are oriented in a certain polarization (i.e. certain atoms

or ions in the crystal have the same distortion direction), and other regions are differently oriented.

Regions with like polarization are called domains, and the boundaries between them are called

domains walls, while c- and c+ domain refers to negative and positive potentials, respectively.

These domains have opposite polarization directions, and thus, the polarization can be depleted,

which is called depolarization [95]. It has been proven that polarization can have an effect on the

photoinduced charges in a photovoltaic solar cell, which tends to promote charge separation [96-

104]. In addition, the photoelectrochemical performance is also affected by polarization materials

[105-109]. Beside strong charge separation, ferroelectric materials have a high adsorption

efficiency due to polar behaviour.

Hybrid ferroelectric semiconductors have a considerable effect on photocatalytic activity. The

photovoltaic effect in a ferroelectric significantly enhances charge separation, which sometimes

exceeds the band gap. Spatial charge separation has been clearly observed on the surface of

ferroelectric BaTiO3. PbO2 and Ag0 have been photodeposited on the selected surfaces where the

domain orientation has a positive or negative end, and thus, Ag+ cations are reduced to Ag on

BaTiO3 where the domain is positive and Pb+ cations are oxidized to PbO2 on BaTiO3 where the

domain is negative[110]. The efficiency increased by a factor of 7.4 when the polarization of

K0.5Na0.5NbO3 via solid solution KxNa1_xNbO3 was increased [111]. The enhanced polarization of

tetragonal BaTiO3 (non-ferroelectric) with respect to cubic BaTiO3 (ferroelectric) results in

enhanced photocatalytic activity. It was confirmed that the reason behind the enhancement of

activity is polarization, which is known to promote charge separation [112]. An interfacing

semiconductor-ferroelectric/ semiconductor, for example, Ag2O−BaTiO3, was previously studied,

and the polarization was maximized by applying ultrasonic waves. As a result, the performance

efficiency was significantly enhanced. Separated redox sites have been confirmed for single

domain PbTiO3 nanoplates: the upper side is the location for reduction, where electrons are

attracted by the positive polarization, and the bottom side is the location for oxidation, where the

holes are attracted to the negative polarization. Thus, PbTiO3 has shown selective photodeposition

for Pt and MnOx (Figure 2-19), and the results demonstrated that selectively photodeposited Pt

Page 42: P-block-based Ferroelectric-photocatalyst compounds

29

features much higher performance in hydrogen production compared with random Pt deposition

[113]. The p-type semiconductor BiFeO3 is considered a promising photocatalyst because its

narrow band gap (~2.4 eV) and ferroelectric properties, while BiFeO3 photoelectrode

demonstrated that higher photocurrent density was observed in downward polarization

(photoelectrode/electrolyte interface) than upward polarization, indicating that the polarization

direction at the contact with electrolyte has a considerable effect on the photoactivity [105].

Figure 2-19 (a) Schematic illustration of single domain PbTiO3 nanoplate

showing the polarization direction; (b) SEM image of selective photodeposition

of Pt and MnOx with schematic illustration in the inset [113].

According to the piezo-phototronic effect, the piezoelectric materials could also behave as

ferroelectric materials and promote charge separation, except that the piezoelectric materials are

activated by external stress such as ultrasound waves or thermal stress. A ZnO-based piezoelectric

has exhibited efficient charge separation on applying external strain. Significant enhancement of

dye degradation was achieved by the piezoelectric photocatalysts Ag2O/ZnO [114], CuS/ZnO

[115], and TiO2/ZnO [116] on applying light and ultrasound waves. Figure 2-20 illustrates the

mechanism of the effect of polarization on charge separation, where the electrons of Ag2O have

been attracted to the positive potential side of BaTiO3, whereas the holes have been attracted to the

negative side of BaTiO3, leading to electron-hole pair separation. Replacing some I5+ by V5+ in IO3

significantly increases the polarization of ferroelectric BiOIO3, which leads to higher charge

separation (Figure 2-21). The enhancement of photocatalytic performance in V-doped BiOIO3 was

10.1 and 21.1 times that of undoped BiOIO3 under visible and UV light. Ferroelectric

photocatalysts based on oxide perovskites (ABO3) are mostly wide-band-gap and have low charge

mobility [117-119]. Similarly, layered structures with spontaneous polarization such as Bi2O2X (X

= SiO3, GeO3) also shows high photocatalytic activity under UV-visible light. The crystal structure

consists a single layer of silicate (or GeO3) sandwiched between bismuth oxide layers. Distortion

is observed in the silicate chain (c-axis), which is twisted by an angle, and consequently, the Bi

ions are slightly distorted, inducing spontaneous polarization perpendicular to the silicate chain,

and thus, 2D polarization (out-of-plane and in-plane polarization) has been observed, so it is

expected that the charge separation is higher in this kind of material.

Page 43: P-block-based Ferroelectric-photocatalyst compounds

30

Figure 2-20 Schematic diagram (a) showing random movement of

photoinduced electron-hole pair; (b) arrangement of photoinduced electron-hole

pairs, with the electrons attracted to the positive surface of BaTiO3, whereas the

holes are attracted to the negative surface of BaTiO3 [120].

Figure 2-21 (a and b) Photodegradation of RhB by BOIO3 samples

under visible and UV light, respectively; (c) Transient photocurrent and

(d) photovoltage spectra for V-doped BiOIO3 and BiOIO3 [121] .

According to theoretical studies, the CB of bismuth silicate is mainly composed of O 2p and Bi

6s orbitals, whereas the top of the valence band consists of Si 3p and O 2p. Consequently, the sp/p

electronic structure means that there is a highly dispersive band structure, leading to high mobility

carriers that facilitate charge transfer to the surface faster than the relaxation time or

recombination time [122]. On the other hand, Ag-based oxides, such as Ag3PO4 [123-126], Ag2O,

Page 44: P-block-based Ferroelectric-photocatalyst compounds

31

and AgCO3 [127-129], have excellent visible light activity with a highly dispersive band structure

similar to those of bismuth oxyhalides. Nevertheless, this photocatalyst has low quantum

efficiency. Previous reports have shown that Ag-based silicates have their own internal electric

field, such as in the case of Ag6Si2O7 and Ag9(SiO4)2NO3, which have been demonstrated

excellent photocatalytic activity, as shown in Figure 2-22 [130].

Figure 2-22 Schematic diagram of the crystal structure of

Ag9(SiO4)2NO3. E is the internal electric field caused by the

polar crystal structure [130].

2.2 Summary and opportunities

The back-reaction of intermediates, a wide band gap and high rate recombination at the surface

of the photocatalyst are major drawbacks to facet photocatalysis. Decreasing the particle size is of

limited help, and the recombination rate will be made to increase, as well the band gap, by

quantum confinement. Heterojunctions (including p-n junctions or phase junctions) have attracted

wide attention due to their ability to absorb a wider spectrum and improve the charge separation.

Heterojunctions still have low efficiency, however, because the separation process occurs only in

the interface and near interface regions, and also, the enhancement is not high, according to review

[83] (except for facet-selective photodeposition [84]). Despite the great efforts previously made to

enhance the photocatalytic activity of these photocatalysts to fulfil the requirements of practical

applications, it is still low. Therefore, to overcome this issue, strategies were adopted in this work

based on the design and engineering of materials with internal electric field properties, such as p-

block materials with ferroelectric polarization or p-block materials with heterojunctions providing

built-in internal electric fields.

2.3 References

1. Fujishima, A. and K. Honda, Electrochemical photolysis of water at a semiconductor electrode.

nature, 1972. 238(5358): p. 37.

2. Chatterjee, D. and S. Dasgupta, Visible light induced photocatalytic degradation of organic

pollutants. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 2005. 6(2–

3): p. 186-205.

3. Kazuhito, H., I. Hiroshi, and F. Akira, TiO2 Photocatalysis: A Historical Overview and Future

Prospects. Japanese Journal of Applied Physics, 2005. 44(12R): p. 8269.

Page 45: P-block-based Ferroelectric-photocatalyst compounds

32

4. Schneider, J., M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi, M. Anpo, and D.W.

Bahnemann, Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chemical

Reviews, 2014. 114(19): p. 9919-9986.

5. Etacheri, V., C. Di Valentin, J. Schneider, D. Bahnemann, and S.C. Pillai, Visible-light

activation of TiO2 photocatalysts: Advances in theory and experiments. Journal of

Photochemistry and Photobiology C: Photochemistry Reviews, 2015. 25: p. 1-29.

6. Asahi, R., T. Morikawa, T. Ohwaki, K. Aoki, and Y. Taga, Visible-light photocatalysis in

nitrogen-doped titanium oxides. science, 2001. 293(5528): p. 269-271.

7. Ohno, T., T. Mitsui, and M. Matsumura, Photocatalytic activity of S-doped TiO2 photocatalyst

under visible light. Chemistry letters, 2003. 32(4): p. 364-365. 8. Iwasaki, M., M. Hara, H. Kawada, H. Tada, and S. Ito, Cobalt ion-doped TiO2 photocatalyst

response to visible light. Journal of Colloid and Interface Science, 2000. 224(1): p. 202-204.

9. Amano, F., E. Ishinaga, and A. Yamakata, Effect of particle size on the photocatalytic activity of

WO3 particles for water oxidation. The Journal of Physical Chemistry C, 2013. 117(44): p.

22584-22590.

10. Nishimoto, S., T. Mano, Y. Kameshima, and M. Miyake, Photocatalytic water treatment over

WO3 under visible light irradiation combined with ozonation. Chemical Physics Letters, 2010.

500(1): p. 86-89.

11. Ashkarran, A., M. Ahadian, and S.M. Ardakani, Synthesis and photocatalytic activity of WO3

nanoparticles prepared by the arc discharge method in deionized water. Nanotechnology, 2008.

19(19): p. 195709. 12. Jiang, H.-Y., J. Liu, K. Cheng, W. Sun, and J. Lin, Enhanced Visible Light Photocatalysis of

Bi2O3 upon Fluorination. The Journal of Physical Chemistry C, 2013. 117(39): p. 20029-20036.

13. Muruganandham, M., R. Amutha, G.-J. Lee, S.-H. Hsieh, J.J. Wu, and M. Sillanpää, Facile

Fabrication of Tunable Bi2O3 Self-Assembly and Its Visible Light Photocatalytic Activity. The

Journal of Physical Chemistry C, 2012. 116(23): p. 12906-12915.

14. Zhang, L., W. Wang, J. Yang, Z. Chen, W. Zhang, L. Zhou, and S. Liu, Sonochemical synthesis

of nanocrystallite Bi2O3 as a visible-light-driven photocatalyst. Applied Catalysis A: General,

2006. 308(Supplement C): p. 105-110.

15. Lang, D., Q. Xiang, G. Qiu, X. Feng, and F. Liu, Effects of crystalline phase and morphology on

the visible light photocatalytic H2-production activity of CdS nanocrystals. Dalton Transactions,

2014. 43(19): p. 7245-7253. 16. Jang, J.S., W. Li, S.H. Oh, and J.S. Lee, Fabrication of CdS/TiO2 nano-bulk composite

photocatalysts for hydrogen production from aqueous H2S solution under visible light. Chemical

Physics Letters, 2006. 425(4): p. 278-282.

17. Zong, X., H. Yan, G. Wu, G. Ma, F. Wen, L. Wang, and C. Li, Enhancement of Photocatalytic

H2 Evolution on CdS by Loading MoS2 as Cocatalyst under Visible Light Irradiation. Journal of

the American Chemical Society, 2008. 130(23): p. 7176-7177.

18. Low, J., B. Cheng, and J. Yu, Surface modification and enhanced photocatalytic CO2 reduction

performance of TiO2: a review. Applied Surface Science, 2017. 392(15): p. 658-686.

19. Pelaez, M., N.T. Nolan, S.C. Pillai, M.K. Seery, P. Falaras, A.G. Kontos, P.S.M. Dunlop, J.W.J.

Hamilton, J.A. Byrne, K. O'Shea, M.H. Entezari, and D.D. Dionysiou, A review on the visible

light active titanium dioxide photocatalysts for environmental applications. Applied Catalysis B:

Environmental, 2012. 125(21): p. 331-349. 20. Linsebigler, A.L., G. Lu, and J.T. Yates Jr, Photocatalysis on TiO2 surfaces: principles,

mechanisms, and selected results. Chemical reviews, 1995. 95(3): p. 735-758.

21. Yan, H., X. Wang, M. Yao, and X. Yao, Band structure design of semiconductors for enhanced

photocatalytic activity: The case of TiO2. Progress in Natural Science: Materials International,

2013. 23(4): p. 402-407.

22. Ullah, R. and J. Dutta, Photocatalytic degradation of organic dyes with manganese-doped ZnO

nanoparticles. Journal of Hazardous Materials, 2008. 156(1): p. 194-200.

23. Lee, K.M., C.W. Lai, K.S. Ngai, and J.C. Juan, Recent developments of zinc oxide based

photocatalyst in water treatment technology: A review. Water Research, 2016. 88(Supplement

C): p. 428-448.

24. Ouyang, S., N. Kikugawa, D. Chen, Z. Zou, and J. Ye, A Systematical Study on Photocatalytic Properties of AgMO2 (M = Al, Ga, In): Effects of Chemical Compositions, Crystal Structures,

and Electronic Structures. The Journal of Physical Chemistry C, 2009. 113(4): p. 1560-1566.

25. Xu, D., B. Cheng, J. Zhang, W. Wang, J. Yu, and W. Ho, Photocatalytic activity of Ag2MO4 (M

= Cr, Mo, W) photocatalysts. Journal of Materials Chemistry A, 2015. 3(40): p. 20153-20166.

Page 46: P-block-based Ferroelectric-photocatalyst compounds

33

26. Wang, Y., R. Zhang, J. Li, L. Li, and S. Lin, First-principles study on transition metal-doped

anatase TiO2. Nanoscale Research Letters, 2014. 9(1): p. 46.

27. Zhang, X., L.-W. Wang, C.-Y. Wang, W.-K. Wang, Y.-L. Chen, Y.-X. Huang, W.-W. Li, Y.-J.

Feng, and H.-Q. Yu, Synthesis of BiOClxBr1−x Nanoplate Solid Solutions as a Robust

Photocatalyst with Tunable Band Structure. Chemistry – A European Journal, 2015. 21(33): p.

11872-11877.

28. Zhang, X., Z. Ai, F. Jia, and L. Zhang, Generalized One-Pot Synthesis, Characterization, and

Photocatalytic Activity of Hierarchical BiOX (X = Cl, Br, I) Nanoplate Microspheres. The

Journal of Physical Chemistry C, 2008. 112(3): p. 747-753.

29. Zhang, W., Q. Zhang, and F. Dong, Visible-Light Photocatalytic Removal of NO in Air over BiOX (X = Cl, Br, I) Single-Crystal Nanoplates Prepared at Room Temperature. Industrial &

Engineering Chemistry Research, 2013. 52(20): p. 6740-6746.

30. Ropero-Vega, J.L., K.L. Rosas-Barrera, J.A. Pedraza-Avella, D.A. Laverde-Cataño, J.E.

Pedraza-Rosas, and M.E. Niño-Gómez, Photophysical and photocatalytic properties of

Bi2MNbO7 (M=Al, In, Ga, Fe) thin films prepared by dip-coating. Materials Science and

Engineering: B, 2010. 174(1): p. 196-199.

31. Wang, L., J. Shang, W. Hao, S. Jiang, S. Huang, T. Wang, Z. Sun, Y. Du, S. Dou, T. Xie, D.

Wang, and J. Wang, A dye-sensitized visible light photocatalyst-Bi24O31Cl10. Sci. Rep., 2014.

4: p. 7384.

32. Shang, J., W. Hao, X. Lv, T. Wang, X. Wang, Y. Du, S. Dou, T. Xie, D. Wang, and J. Wang,

Bismuth Oxybromide with Reasonable Photocatalytic Reduction Activity under Visible Light. ACS Catalysis, 2014. 4(3): p. 954-961.

33. Marchelek, M., M. Diak, M. Kozak, A. Zaleska-Medynska, and E. Grabowska, Some Unitary,

Binary, and Ternary Non-TiO2 Photocatalysts, in Semiconductor Photocatalysis - Materials,

Mechanisms and Applications, W. Cao, Editor. 2016, InTech: Rijeka. p. Ch. 09.

34. Li, F.-t., Q. Wang, X.-j. Wang, B. Li, Y.-j. Hao, R.-h. Liu, and D.-s. Zhao, In-situ one-step

synthesis of novel BiOCl/Bi24O31Cl10 heterojunctions via self-combustion of ionic liquid with

enhanced visible-light photocatalytic activities. Applied Catalysis B: Environmental, 2014.

150(Supplement C): p. 574-584.

35. Jin, X., L. Ye, H. Wang, Y. Su, H. Xie, Z. Zhong, and H. Zhang, Bismuth-rich strategy induced

photocatalytic molecular oxygen activation properties of bismuth oxyhalogen: The case of

Bi24O31Cl10. Applied Catalysis B: Environmental, 2015. 165(Supplement C): p. 668-675. 36. Bai, Y., T. Chen, P. Wang, L. Wang, and L. Ye, Bismuth-rich Bi4O5X2 (X=Br, and I) nanosheets

with dominant {101} facets exposure for photocatalytic H2 evolution. Chemical Engineering

Journal, 2016. 304(Supplement C): p. 454-460.

37. Wang, J., Y. Yu, and L. Zhang, Highly efficient photocatalytic removal of sodium

pentachlorophenate with Bi3O4Br under visible light. Applied Catalysis B: Environmental, 2013.

136(5): p. 112-121.

38. Jiang, R., B. Li, C. Fang, and J. Wang, Metal/Semiconductor Hybrid Nanostructures for

Plasmon‐Enhanced Applications. Advanced Materials, 2014. 26(31): p. 5274-5309.

39. Kale, M.J., T. Avanesian, and P. Christopher, Direct Photocatalysis by Plasmonic

Nanostructures. ACS Catalysis, 2014. 4(1): p. 116-128.

40. Khan, M.R., T.W. Chuan, A. Yousuf, M.N.K. Chowdhury, and C.K. Cheng, Schottky barrier

and surface plasmonic resonance phenomena towards the photocatalytic reaction: study of their mechanisms to enhance photocatalytic activity. Catalysis Science & Technology, 2015. 5(5): p.

2522-2531.

41. Yu, C., F. Cao, G. Li, R. Wei, J.C. Yu, R. Jin, Q. Fan, and C. Wang, Novel noble metal (Rh, Pd,

Pt)/BiOX(Cl, Br, I) composite photocatalysts with enhanced photocatalytic performance in dye

degradation. Separation and Purification Technology, 2013. 120: p. 110-122.

42. Xiao, X., L. Ge, C. Han, Y. Li, Z. Zhao, Y. Xin, S. Fang, L. Wu, and P. Qiu, A facile way to

synthesize Ag@AgBr cubic cages with efficient visible-light-induced photocatalytic activity.

Applied Catalysis B: Environmental, 2015. 163: p. 564-572.

43. Hu, H., Z. Jiao, T. Wang, J. Ye, G. Lu, and Y. Bi, Enhanced photocatalytic activity of

Ag/Ag3PO4 coaxial hetero-nanowires. Journal of Materials Chemistry A, 2013. 1(36): p. 10612-

10616. 44. Yan, T., H. Zhang, Y. Liu, W. Guan, J. Long, W. Li, and J. You, Fabrication of robust

M/Ag3PO4 (M = Pt, Pd, Au) Schottky-type heterostructures for improved visible-light

photocatalysis. RSC Advances, 2014. 4(70): p. 37220-37230.

Page 47: P-block-based Ferroelectric-photocatalyst compounds

34

45. Bai, S., X. Li, Q. Kong, R. Long, C. Wang, J. Jiang, and Y. Xiong, Toward Enhanced

Photocatalytic Oxygen Evolution: Synergetic Utilization of Plasmonic Effect and Schottky

Junction via Interfacing Facet Selection. Advanced Materials, 2015. 27(22): p. 3444-3452.

46. Reza Gholipour, M., C.-T. Dinh, F. Beland, and T.-O. Do, Nanocomposite heterojunctions as

sunlight-driven photocatalysts for hydrogen production from water splitting. Nanoscale, 2015.

7(18): p. 8187-8208.

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

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

Performance under LED Light Irradiation. ACS Applied Materials & Interfaces, 2015. 7(17): p.

9023-9030. 48. Wang, C., C. Shao, X. Zhang, and Y. Liu, SnO2 Nanostructures-TiO2 Nanofibers

Heterostructures: Controlled Fabrication and High Photocatalytic Properties. Inorganic

Chemistry, 2009. 48(15): p. 7261-7268.

49. Yu, J.C., L. Wu, J. Lin, P. Li, and Q. Li, Microemulsion-mediated solvothermal synthesis of

nanosized CdS-sensitized TiO2 crystalline photocatalyst. Chemical Communications, 2003.

0(13): p. 1552-1553.

50. Li, F.-t., Q. Wang, J. Ran, Y.-j. Hao, X.-j. Wang, D. Zhao, and S.Z. Qiao, Ionic liquid self-

combustion synthesis of BiOBr/Bi24O31Br10 heterojunctions with exceptional visible-light

photocatalytic performances. Nanoscale, 2015. 7(3): p. 1116-1126.

51. Dai, Z., F. Qin, H. Zhao, F. Tian, Y. Liu, and R. Chen, Time-dependent evolution of the

Bi3.64Mo0.36O6.55/Bi2MoO6 heterostructure for enhanced photocatalytic activity via the interfacial hole migration. Nanoscale, 2015. 7(28): p. 11991-11999.

52. Ge, M., Y. Li, L. Liu, Z. Zhou, and W. Chen, Bi2O3−Bi2WO6 Composite Microspheres:

Hydrothermal Synthesis and Photocatalytic Performances. The Journal of Physical Chemistry C,

2011. 115(13): p. 5220-5225.

53. Cao, J., B. Xu, B. Luo, H. Lin, and S. Chen, Novel BiOI/BiOBr heterojunction photocatalysts

with enhanced visible light photocatalytic properties. Catalysis Communications, 2011. 13(1): p.

63-68.

54. Huang, H., S. Wang, N. Tian, and Y. Zhang, A one-step hydrothermal preparation strategy for

layered BiIO4/Bi2WO6 heterojunctions with enhanced visible light photocatalytic activities. RSC

Advances, 2014. 4(11): p. 5561-5567.

55. Cui, Y., Q. Jia, H. Li, J. Han, L. Zhu, S. Li, Y. Zou, and J. Yang, Photocatalytic activities of Bi2S3/BiOBr nanocomposites synthesized by a facile hydrothermal process. Applied Surface

Science, 2014. 290: p. 233-239.

56. Sun, W.-T., Y. Yu, H.-Y. Pan, X.-F. Gao, Q. Chen, and L.-M. Peng, CdS quantum dots

sensitized TiO2 nanotube-array photoelectrodes. Journal of the American Chemical Society,

2008. 130(4): p. 1124-1125.

57. Li, G., L. Wu, F. Li, P. Xu, D. Zhang, and H. Li, Photoelectrocatalytic degradation of organic

pollutants via a CdS quantum dots enhanced TiO2 nanotube array electrode under visible light

irradiation. nanoscale, 2013. 5(5): p. 2118-2125.

58. Zhang, Z., J. Huang, M. Zhang, Q. Yuan, and B. Dong, Ultrathin hexagonal SnS2 nanosheets

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

photocatalytic activity. Applied Catalysis B: Environmental, 2015. 163(Supplement C): p. 298-

305. 59. Yuan, L., B. Weng, J.C. Colmenares, Y. Sun, and Y.-J. Xu, Multichannel Charge Transfer and

Mechanistic Insight in Metal Decorated 2D–2D Bi2WO6–TiO2 Cascade with Enhanced

Photocatalytic Performance. Small, 2017. 13(48): p. 1702253.

60. Xu, H., J. Zhu, Y. Song, W. Zhao, Y. Xu, Y. Song, H. Ji, and H. Li, Ion-exchange preparation

for visible-light-driven photocatalyst AgBr/Ag2CO3 and its photocatalytic activity. RSC

Advances, 2014. 4(18): p. 9139-9147.

61. Yu, C., G. Li, S. Kumar, K. Yang, and R. Jin, Phase Transformation Synthesis of Novel

Ag2O/Ag2CO3 Heterostructures with High Visible Light Efficiency in Photocatalytic

Degradation of Pollutants. Advanced Materials, 2014. 26(6): p. 892-898.

62. Liang, N., M. Wang, L. Jin, S. Huang, W. Chen, M. Xu, Q. He, J. Zai, N. Fang, and X. Qian,

Highly Efficient Ag2O/Bi2O2CO3 p-n Heterojunction Photocatalysts with Improved Visible-Light Responsive Activity. ACS Applied Materials & Interfaces, 2014. 6(14): p. 11698-11705.

63. Liang, N., J. Zai, M. Xu, Q. Zhu, X. Wei, and X. Qian, Novel Bi2S3/Bi2O2CO3 heterojunction

photocatalysts with enhanced visible light responsive activity and wastewater treatment. Journal

of Materials Chemistry A, 2014. 2(12): p. 4208-4216.

Page 48: P-block-based Ferroelectric-photocatalyst compounds

35

64. Wang, P.Q., Y. Bai, P.Y. Luo, and J.Y. Liu, Ag2O/Ag3PO4 heterostructures: highly efficient and

stable visible-light-induced photocatalyst for degradation of methyl orange and phenol. IET

Micro & Nano Letters, 2013. 8(7): p. 340-344.

65. Li, L., P.A. Salvador, and G.S. Rohrer, Photocatalysts with internal electric fields. Nanoscale,

2014. 6(1): p. 24-42.

66. Xu, H., C. Wang, Y. Song, J. Zhu, Y. Xu, J. Yan, Y. Song, and H. Li, CNT/Ag3PO4 composites

with highly enhanced visible light photocatalytic activity and stability. Chemical Engineering

Journal, 2014. 241: p. 35-42.

67. Liu, J.J., X.L. Fu, S.F. Chen, and Y.F. Zhu, Electronic structure and optical properties of

Ag3PO4 photocatalyst calculated by hybrid density functional method. Applied Physics Letters, 2011. 99(19): p. 191903.

68. Xiang, Q., D. Lang, T. Shen, and F. Liu, Graphene-modified nanosized Ag3PO4 photocatalysts

for enhanced visible-light photocatalytic activity and stability. Applied Catalysis B:

Environmental, 2015. 162: p. 196-203.

69. Cui, C., Y. Wang, D. Liang, W. Cui, H. Hu, B. Lu, S. Xu, X. Li, C. Wang, and Y. Yang, Photo-

assisted synthesis of Ag3PO4/reduced graphene oxide/Ag heterostructure photocatalyst with

enhanced photocatalytic activity and stability under visible light. Applied Catalysis B:

Environmental, 2014. 158: p. 150-160.

70. Yang, X., H. Cui, Y. Li, J. Qin, R. Zhang, and H. Tang, Fabrication of Ag3PO4-Graphene

Composites with Highly Efficient and Stable Visible Light Photocatalytic Performance. ACS

Catalysis, 2013. 3(3): p. 363-369. 71. Liu, Y., L. Fang, H. Lu, L. Liu, H. Wang, and C. Hu, Highly efficient and stable Ag/Ag3PO4

plasmonic photocatalyst in visible light. Catalysis Communications, 2012. 17: p. 200-204.

72. Vinoth, R., S.G. Babu, R. Ramachandran, and B. Neppolian, Bismuth oxyiodide incorporated

reduced graphene oxide nanocomposite material as an efficient photocatalyst for visible light

assisted degradation of organic pollutants. Applied Surface Science, 2017. 418: p. 163-170.

73. Gao, B., G.Z. Chen, and G. Li Puma, Carbon nanotubes/titanium dioxide (CNTs/TiO2)

nanocomposites prepared by conventional and novel surfactant wrapping sol–gel methods

exhibiting enhanced photocatalytic activity. Applied Catalysis B: Environmental, 2009. 89(3): p.

503-509.

74. Lee, W.J., J.M. Lee, S.T. Kochuveedu, T.H. Han, H.Y. Jeong, M. Park, J.M. Yun, J. Kwon, K.

No, D.H. Kim, and S.O. Kim, Biomineralized N-Doped CNT/TiO2 Core/Shell Nanowires for Visible Light Photocatalysis. ACS Nano, 2012. 6(1): p. 935-943.

75. Peng, T., K. Li, P. Zeng, Q. Zhang, and X. Zhang, Enhanced Photocatalytic Hydrogen

Production over Graphene Oxide–Cadmium Sulfide Nanocomposite under Visible Light

Irradiation. The Journal of Physical Chemistry C, 2012. 116(43): p. 22720-22726.

76. Zhang, S., H. Gao, X. Liu, Y. Huang, X. Xu, N.S. Alharbi, T. Hayat, and J. Li, Hybrid 0D–2D

Nanoheterostructures: In Situ Growth of Amorphous Silver Silicates Dots on g-C3N4

Nanosheets for Full-Spectrum Photocatalysis. ACS Applied Materials & Interfaces, 2016. 8(51):

p. 35138-35149.

77. Fang, Z., Y. Wang, J. Song, Y. Sun, J. Zhou, R. Xu, and H. Duan, Immobilizing CdS quantum

dots and dendritic Pt nanocrystals on thiolated graphene nanosheets toward highly efficient

photocatalytic H2 evolution. Nanoscale, 2013. 5(20): p. 9830-9838.

78. Maeda, K., M. Eguchi, W.J. Youngblood, and T.E. Mallouk, Niobium Oxide Nanoscrolls as Building Blocks for Dye-Sensitized Hydrogen Production from Water under Visible Light

Irradiation. Chemistry of Materials, 2008. 20(21): p. 6770-6778.

79. Lopes, O.F., E.C. Paris, and C. Ribeiro, Synthesis of Nb2O5 nanoparticles through the oxidant

peroxide method applied to organic pollutant photodegradation: A mechanistic study. Applied

Catalysis B: Environmental, 2014. 144: p. 800-808.

80. Zhang, X., T. Peng, and S. Song, Recent advances in dye-sensitized semiconductor systems for

photocatalytic hydrogen production. Journal of Materials Chemistry A, 2016. 4(7): p. 2365-

2402.

81. Wang, Z. and X. Lang, Visible light photocatalysis of dye-sensitized TiO2: The selective aerobic

oxidation of amines to imines. Applied Catalysis B: Environmental, 2018. 224: p. 404-409.

82. Li, X., J.-L. Shi, H. Hao, and X. Lang, Visible light-induced selective oxidation of alcohols with air by dye-sensitized TiO2 photocatalysis. Applied Catalysis B: Environmental, 2018. 232(15): p.

260-267.

83. Shamim, N. and V.K. Sharma, eds. Sustainable Nanotechnology and the Environment: Advances

and Achievements. ACS Symposium Series. Vol. 1124. 2013, American Chemical Society. 0.

Page 49: P-block-based Ferroelectric-photocatalyst compounds

36

84. Xiao, X., C. Liu, R. Hu, X. Zuo, J. Nan, L. Li, and L. Wang, Oxygen-rich bismuth oxyhalides:

generalized one-pot synthesis, band structures and visible-light photocatalytic properties.

Journal of Materials Chemistry, 2012. 22(43): p. 22840-22843.

85. Zhang, R., Y. Dai, Z. Lou, Z. Li, Z. Wang, Y. Yang, X. Qin, X. Zhang, and B. Huang, Layered

photocatalyst Bi2O2[BO2(OH)] nanosheets with internal polar field enhanced photocatalytic

activity. CrystEngComm, 2014. 16(23): p. 4931-4934.

86. Mi, Y., M. Zhou, L. Wen, H. Zhao, and Y. Lei, A highly efficient visible-light driven

photocatalyst: two dimensional square-like bismuth oxyiodine nanosheets. Dalton Transactions,

2014. 43(25): p. 9549-9556.

87. Jiang, J., K. Zhao, X. Xiao, and L. Zhang, Synthesis and Facet-Dependent Photoreactivity of BiOCl Single-Crystalline Nanosheets. Journal of the American Chemical Society, 2012. 134(10):

p. 4473-4476.

88. Yang, H.G., C.H. Sun, S.Z. Qiao, J. Zou, G. Liu, S.C. Smith, H.M. Cheng, and G.Q. Lu, Anatase

TiO2 single crystals with a large percentage of reactive facets. Nature, 2008. 453(7195): p. 638-

641.

89. Wang, J., F. Teng, M. Chen, J. Xu, Y. Song, and X. Zhou, Facile synthesis of novel Ag3PO4

tetrapods and the {110} facets-dominated photocatalytic activity. CrystEngComm, 2013. 15(1):

p. 39-42.

90. Pan, M., H. Zhang, G. Gao, L. Liu, and W. Chen, Facet-Dependent Catalytic Activity of

Nanosheet-Assembled Bismuth Oxyiodide Microspheres in Degradation of Bisphenol A.

Environmental Science & Technology, 2015. 49(10): p. 6240-6248. 91. Li, Q., X. Zhao, J. Yang, C.-J. Jia, Z. Jin, and W. Fan, Exploring the effects of nanocrystal facet

orientations in g-C3N4/BiOCl heterostructures on photocatalytic performance. Nanoscale, 2015.

7(45): p. 18971-18983.

92. Li, J., L. Cai, J. Shang, Y. Yu, and L. Zhang, Giant Enhancement of Internal Electric Field

Boosting Bulk Charge Separation for Photocatalysis. Advanced Materials, 2016. 28(21): p.

4059-4064.

93. Sun, L., L. Xiang, X. Zhao, C.-J. Jia, J. Yang, Z. Jin, X. Cheng, and W. Fan, Enhanced Visible-

Light Photocatalytic Activity of BiOI/BiOCl Heterojunctions: Key Role of Crystal Facet

Combination. ACS Catalysis, 2015. 5(6): p. 3540-3551.

94. Wang, L., J. Ge, A. Wang, M. Deng, X. Wang, S. Bai, R. Li, J. Jiang, Q. Zhang, Y. Luo, and Y.

Xiong, Designing p-Type Semiconductor–Metal Hybrid Structures for Improved Photocatalysis. Angewandte Chemie International Edition, 2014. 53(20): p. 5107-5111.

95. Fong, D.D., A.M. Kolpak, J.A. Eastman, S.K. Streiffer, P.H. Fuoss, G.B. Stephenson, C.

Thompson, D.M. Kim, K.J. Choi, C.B. Eom, I. Grinberg, and A.M. Rappe, Stabilization of

Monodomain Polarization in Ultrathin PbTiO3 Films. Physical Review Letters, 2006. 96(12): p.

127601.

96. Yao, K., B.K. Gan, M. Chen, and S. Shannigrahi, Large photo-induced voltage in a ferroelectric

thin film with in-plane polarization. Applied Physics Letters, 2005. 87(21): p. 212906.

97. Matsuo, H., Y. Noguchi, and M. Miyayama, Gap-state engineering of visible-light-active

ferroelectrics for photovoltaic applications. Nature Communications, 2017. 8(1): p. 207.

98. Yang, S.Y., J. Seidel, S.J. Byrnes, P. Shafer, C.H. Yang, M.D. Rossell, P. Yu, Y.H. Chu, J.F.

Scott, J.W. Ager Iii, L.W. Martin, and R. Ramesh, Above-bandgap voltages from ferroelectric

photovoltaic devices. Nature Nanotechnology, 2010. 5: p. 143. 99. Goldstein, B. and L. Pensak, High‐Voltage Photovoltaic Effect. Journal of Applied Physics,

1959. 30(2): p. 155-161.

100. Shunsuke, M., W. Takayuki, S. Tatsuya, M. Takanori, and M. Kaoru, Effects of poling

termination and aging process on piezoelectric properties of Mn-doped BaTi0.96Zr0.04O3

ceramics. Japanese Journal of Applied Physics, 2015. 54(10S): p. 10ND05.

101. Inoue, R., S. Ishikawa, R. Imura, Y. Kitanaka, T. Oguchi, Y. Noguchi, and M. Miyayama, Giant

photovoltaic effect of ferroelectric domain walls in perovskite single crystals. Scientific Reports,

2015. 5: p. 14741.

102. Kenji, U., M. Yuichi, and N. Shoichiro, High-Voltage Photovoltaic Effect in PbTiO3 -Based

Ceramics. Japanese Journal of Applied Physics, 1982. 21(12R): p. 1671.

103. Matsuo, H., Y. Kitanaka, R. Inoue, Y. Noguchi, M. Miyayama, T. Kiguchi, and T.J. Konno, Bulk and domain-wall effects in ferroelectric photovoltaics. Physical Review B, 2016. 94(21): p.

214111.

104. Zhang, G., H. Wu, G. Li, Q. Huang, C. Yang, F. Huang, F. Liao, and J. Lin, New high Tc

multiferroics KBiFe2O5 with narrow band gap and promising photovoltaic effect. Scientific

Reports, 2013. 3: p. 1265.

Page 50: P-block-based Ferroelectric-photocatalyst compounds

37

105. Liu, Q., Y. Zhou, L. You, J. Wang, M. Shen, and L. Fang, Enhanced ferroelectric

photoelectrochemical properties of polycrystalline BiFeO3 film by decorating with Ag

nanoparticles. Applied Physics Letters, 2016. 108(2): p. 022902.

106. Yang, W., Y. Yu, M.B. Starr, X. Yin, Z. Li, A. Kvit, S. Wang, P. Zhao, and X. Wang,

Ferroelectric Polarization-Enhanced Photoelectrochemical Water Splitting in TiO2–BaTiO3

Core–Shell Nanowire Photoanodes. Nano Letters, 2015. 15(11): p. 7574-7580.

107. Li, S., J. Zhang, B.-P. Zhang, W. Huang, C. Harnagea, R. Nechache, L. Zhu, S. Zhang, Y.-H.

Lin, L. Ni, Y.-H. Sang, H. Liu, and F. Rosei, Manipulation of charge transfer in vertically

aligned epitaxial ferroelectric KNbO3 nanowire array photoelectrodes. Nano Energy, 2017.

35(Supplement C): p. 92-100. 108. Xie, J., C. Guo, P. Yang, X. Wang, D. Liu, and C.M. Li, Bi-functional ferroelectric BiFeO3

passivated BiVO4 photoanode for efficient and stable solar water oxidation. Nano Energy, 2017.

31(Supplement C): p. 28-36.

109. Singh, S. and N. Khare, Coupling of piezoelectric, semiconducting and photoexcitation

properties in NaNbO3 nanostructures for controlling electrical transport: Realizing an efficient

piezo-photoanode and piezo-photocatalyst. Nano Energy, 2017. 38(Supplement C): p. 335-341.

110. Giocondi, J.L. and G.S. Rohrer, Spatial Separation of Photochemical Oxidation and Reduction

Reactions on the Surface of Ferroelectric BaTiO3. The Journal of Physical Chemistry B, 2001.

105(35): p. 8275-8277.

111. Park, S., C.W. Lee, M.-G. Kang, S. Kim, H.J. Kim, J.E. Kwon, S.Y. Park, C.-Y. Kang, K.S.

Hong, and K.T. Nam, A ferroelectric photocatalyst for enhancing hydrogen evolution: polarized particulate suspension. Physical Chemistry Chemical Physics, 2014. 16(22): p. 10408-10413.

112. Cui, Y., J. Briscoe, and S. Dunn, Effect of Ferroelectricity on Solar-Light-Driven Photocatalytic

Activity of BaTiO3—Influence on the Carrier Separation and Stern Layer Formation. Chemistry

of Materials, 2013. 25(21): p. 4215-4223.

113. Zhen, C., J.C. Yu, G. Liu, and H.-M. Cheng, Selective deposition of redox co-catalyst(s) to

improve the photocatalytic activity of single-domain ferroelectric PbTiO3 nanoplates. Chemical

Communications, 2014. 50(72): p. 10416-10419.

114. Sun, C., Y. Fu, Q. Wang, L. Xing, B. Liu, and X. Xue, Ultrafast piezo-photocatalytic

degradation of organic pollutions by Ag2O/tetrapod-ZnO nanostructures under ultrasonic/UV

exposure. RSC Advances, 2016. 6(90): p. 87446-87453.

115. Hong, D., W. Zang, X. Guo, Y. Fu, H. He, J. Sun, L. Xing, B. Liu, and X. Xue, High Piezo-photocatalytic Efficiency of CuS/ZnO Nanowires Using Both Solar and Mechanical Energy for

Degrading Organic Dye. ACS Applied Materials & Interfaces, 2016. 8(33): p. 21302-21314.

116. Wang, L., S. Liu, Z. Wang, Y. Zhou, Y. Qin, and Z.L. Wang, Piezotronic Effect Enhanced

Photocatalysis in Strained Anisotropic ZnO/TiO2 Nanoplatelets via Thermal Stress. ACS Nano,

2016. 10(2): p. 2636-2643.

117. Grinberg, I., D.V. West, M. Torres, G. Gou, D.M. Stein, L. Wu, G. Chen, E.M. Gallo, A.R.

Akbashev, P.K. Davies, J.E. Spanier, and A.M. Rappe, Perovskite oxides for visible-light-

absorbing ferroelectric and photovoltaic materials. Nature, 2013. 503(7477): p. 509-512.

118. Mu, L., Y. Zhao, A. Li, S. Wang, Z. Wang, J. Yang, Y. Wang, T. Liu, R. Chen, J. Zhu, F. Fan,

R. Li, and C. Li, Enhancing charge separation on high symmetry SrTiO3 exposed with

anisotropic facets for photocatalytic water splitting. Energy & Environmental Science, 2016.

9(7): p. 2463-2469. 119. Wang, R., Y. Zhu, Y. Qiu, C.-F. Leung, J. He, G. Liu, and T.-C. Lau, Synthesis of nitrogen-

doped KNbO3 nanocubes with high photocatalytic activity for water splitting and degradation of

organic pollutants under visible light. Chemical Engineering Journal, 2013. 226: p. 123-130.

120. Xie, H., F. Li, H. Xi, and D. Zhou, Microwave Dielectric Properties of Sol-Gel Processed

Bi4Si3O12 Ceramics and Single Crystal. Transactions of the Indian Ceramic Society, 2015. 74(2):

p. 83-85.

121. Huang, H., S. Tu, C. Zeng, T. Zhang, A.H. Reshak, and Y. Zhang, Macroscopic Polarization

Enhancement Promoting Photo‐ and Piezoelectric‐Induced Charge Separation and Molecular

Oxygen Activation. Angewandte Chemie International Edition, 2017. 56(39): p. 11860-11864.

122. Shang, J., W. Hao, X. Lv, T. Wang, X. Wang, Y. Du, S. Dou, T. Xie, D. Wang, and J. Wang,

Bismuth Oxybromide with Reasonable Photocatalytic Reduction Activity under Visible Light. ACS Catal., 2014. 4(3): p. 954-961.

123. Teng, W., X. Li, Q. Zhao, J. Zhao, and D. Zhang, In situ capture of active species and oxidation

mechanism of RhB and MB dyes over sunlight-driven Ag/Ag3PO4 plasmonic nanocatalyst.

Applied Catalysis B: Environmental, 2012. 125: p. 538-545.

Page 51: P-block-based Ferroelectric-photocatalyst compounds

38

124. Zhou, C., Y. Zhao, J. Cao, H. Lin, and S. Chen, Partial oxidation controlled activity

regeneration of used Ag3PO4 photocatalyst via removing the in situ surface metallic silver.

Applied Surface Science, 2015. 351: p. 33-39.

125. Wan, J., E. Liu, J. Fan, X. Hu, L. Sun, C. Tang, Y. Yin, H. Li, and Y. Hu, In-situ synthesis of

plasmonic Ag/Ag3PO4 tetrahedron with exposed {111} facets for high visible-light

photocatalytic activity and stability. Ceramics International, 2015. 41(5): p. 6933-6940.

126. Yi, Z., J. Ye, N. Kikugawa, T. Kako, S. Ouyang, H. Stuart-Williams, H. Yang, J. Cao, W. Luo,

Z. Li, Y. Liu, and R.L. Withers, An orthophosphate semiconductor with photooxidation

properties under visible-light irradiation. Nat Mater, 2010. 9(7): p. 559-564.

127. Yan, H., X. Wang, M. Yao, and X. Yao, Band structure design of semiconductors for enhanced photocatalytic activity: The case of TiO2. Prog. Nat. Sci., 2013. 23(4): p. 402-407.

128. Xu, C., Y. Liu, B. Huang, H. Li, X. Qin, X. Zhang, and Y. Dai, Preparation, characterization,

and photocatalytic properties of silver carbonate. Applied Surface Science, 2011. 257(20): p.

8732-8736.

129. Wang, Y., P. Ren, C. Feng, X. Zheng, Z. Wang, and D. Li, Photocatalytic behavior and photo-

corrosion of visible-light-active silver carbonate/titanium dioxide. Materials Letters, 2014. 115:

p. 85-88.

130. Zhu, X., Z. Wang, B. Huang, W. Wei, Y. Dai, X. Zhang, and X. Qin, Synthesis of Ag9(SiO4)2NO3

through a reactive flux method and its visible-light photocatalytic performances. APL Materials,

2015. 3(10): p. 104413.

Page 52: P-block-based Ferroelectric-photocatalyst compounds

39

Chapter 3

Experimental techniques

This chapter presents details about materials and instruments that were used in this project, along

with a discussion of the advanced techniques used to evaluate the photocatalyst. These instruments

used in this project are located in the Institute for Superconducting and Electronic materials

(ISEM), the Electron Microscopy Centre (EMC), and the Intelligent Polymer Research Institute

(IPRI) in the University of Wollongong (Australia), and also the Department of Physics and Key

Laboratory of Micro-nano Measurement, Manipulation and Physics, Ministry of Education

(MOE), Beihang University, Beijing (China).

3.1 Materials:

Table 3-1 all the general chemicals and reagents used in this work.

Chemicals Molecular formula

Molecular

Weight

(g mol-1)

Purity (%) Supplier

Silver Nitrate AgNO3 169.8731 99 Sigma-Aldrich

Sodium Silicate Na2SiO3 122.0632 99 Sigma-Aldrich

Titanium Oxide TiO2 Sigma-Aldrich

Bismuth Nitrate Bi(NO3)3·5H2O 485.07 99 Sigma-Aldrich

Cetyltrimethylammon

ium Bromide (CTAB) C19H42BrN 364.45 99 Sigma-Aldrich

Polyvinylpyrrolidone

(PVP) (C6H9NO)n 30k 99 Sigma-Aldrich

Mannitol C6H14O6 182.172 99 Sigma-Aldrich

Distilled

Water

H2O 18.01528 99 Sigma-Aldrich

Sodium Hydroxide NaOH 39.997 99 Sigma-Aldrich

Ethanol C2H6O 46.06844 97 Sigma-Aldrich

Titanium Butoxide Ti(OBu)₄ 340.32 99 Sigma-Aldrich

Sodium Iodide NaI 149.89 99 Sigma-Aldrich

Page 53: P-block-based Ferroelectric-photocatalyst compounds

40

Dye solutions:

Table 3-2 all the dyes used in this work.

Dyes Molecular

formula

Molecular Weight (g.

mol-1) Supplier

Rhodamine B (RhB) C28H31ClN2O3 479.02 Sigma-Aldrich

Methyl Orange (MO) C14H14N3NaO3S 327.33 Sigma-Aldrich

Methylene Blue (MB) C16H18ClN3S 319.851 Sigma-Aldrich

Pollutants components:

Component Molecular Molecular Weight (g.

mol-1)

Supplier

Phenol C6H6O 94.11 Sigma-Aldrich

3.2 Synthesis

The synthesis of all the photocatalysts in this work was conducted by three different methods,

the details of which are below:

3.2.1 Solid ion-exchange method

Normally, two precursors that need to react should be dissolved in solution (water or some

organic solution) and the ions exchanges between the two materials are produced precipitating

non-soluble materials from the solution, which should be crystallized. In some special cases, it is

hard to produce crystalline materials by this method, such as in the case of silver silicate

compounds, and solid state reaction under high oxygen pressure is very expensive [1]. Thus, it is

more attractive to mix the precursors in mortar and press the mixture into pellets at high pressure,

and then the ion exchange will proceed at a specific temperature and over a specific time without

solvents [2, 3]. The product is then washed several times with water and ethanol to remove

residual materials. For more details, see Chapter 4. This method was used to prepare Ag10Si4O13.

3.2.2 Hydrothermal Method

One of the most popular methods to synthesize uniform nanomaterials with high crystallinity is

the hydrothermal method. Control of the morphology of materials, in particular, for application in

photocatalysis, to produce nanoparticles, nanosheets, nanowires, etc., can be accomplished by the

hydrothermal method [3, 4]. This method is a typical wet-chemical procedure. In this method, the

reactant should be soluble in solvents such as water or any organic solvent under base or acidic

environmental conditions. At high temperatures and pressures, the dissolved materials start to

react, nucleation and crystal growth results in the formation of new materials [4]. In the basic

procedure, the dissolved materials after mixing by stirring are poured into Teflon-lined stainless

steel autoclaves, sealed, and then heated in an oven at a certain temperature and for a certain time,

Page 54: P-block-based Ferroelectric-photocatalyst compounds

41

and the mixtures is allowed to cool to room temperature. The precipitate is washed several times

with distilled water and ethanol to remove residual materials. The advantage of this method is that

thin nanosheets of layered structure materials can be prepared which are difficult to prepare by any

other method, and it is also possible to prepare heterojunctions and to dope materials. The

preparation details are reported in Chapter 5. . This method was used to prepare

Bi2SiO5/Bi4Si3O12.

3.2.3 Molten salt synthesis

Molten salt synthesis is one of the most common methods to prepare complex oxide materials

such as ceramic materials; this method features enhanced rates of solid state reaction and lower

temperature reaction. It is considered that molten salts are good solvents at high temperature.

There are a variety of melting temperatures, and different salts can also be mixed to get an

appropriate melting temperature. Basically, by this method, the reactants and salts are mixed in a

mortar and then heated in furnace to the appropriate molten salt temperature. At this temperature,

the mixture start to react, leading to nucleation and crystal growth. It is possible to prepare

nanostructures with controllable shapes and sizes, and to prepare large amounts by controlling the

temperature, time, and cooling rate as well as solubility of salts [5, 6]. Due to these advantages,

this method was used in this work to prepare thin plates of ternary oxides, which is difficult to

synthesize by conventional methods. These ternary oxides are especially important for preparing

ferroelectric materials which need high temperature[7]. The synthesis of these oxides in different

morphologies, and nanoplates in particular, is possible by this method [8]. Herein, preparation

details are reported in Chapter 6. .This method was used to prepare Bi4Ti3O12.

3.2.4 Thin film synthesis

There are many methods to prepare thin film. Thin film synthesis is important for studying the

electrochemical properties of a photocatalyst. In this project, the thin film was prepared via the

dip-coating method on indium tin oxide (ITO) or fluorine doped tin oxide (FTO) and had an area

of 1 cm2. The substrates should be previously cleaned with ethanol under sonication for 10 min.

The powder was mixed with binder material (such as PVA) and drop it on substrate drop by drop

by pipette and then dried at a specific temperature and for a given time, which provided

mechanical stability.

3.3 Characterization: Equipment, instruments, and tools

3.3.1 Scanning electron microscopy / Energy dispersive X-ray spectroscopy

The morphology and material compositions of the samples were investigated and confirmed by

the field emission scanning electron microscopy / energy dispersive X-ray spectroscopy

(FESEM/EDS). The sample surface is scanned by an electron beam generated by the electron gun

at the top of column, and the electron beam is vertically focused on a small area of the sample

surface by a magnetic lens. The column is kept under vacuum, and a high acceleration voltage is

used to avoid scattering electrons. The electron beam generates secondary electrons after colliding

with the sample surface. These secondary electrons are detected by an electron detector. The

elements in the sample are also detected by X-rays emitted from the sample surface, which can be

measured by an energy-dispersive spectrometer. The SEM produces high resolution images and

Page 55: P-block-based Ferroelectric-photocatalyst compounds

42

very fine details within an area less than 2 nm2. In this work, the FESEM (JEOL-7500) was used

throughout the project (5 kV, 10 μA, 8 mm work distance). Before testing, the powdered

photocatalyst was sprayed on carbon tape, which was previously covered. The tape is put in an

SEM holder, and then air is blown to get a uniform dispersion of particles. Finally, the surface is

coated with a 15 nm layer of Pt to make the powder surface conducting. The magnifications were

different, depending on sample size. For EDS measurements, the settings were different. The

accelerating voltage was 20 kV with the maximum emission current 20 μA and the work distance

10 mm. Many points on each sample surface were measured or mapping scanning of surface.

3.3.2 Transmission electron microscopy (TEM)/EDS

The TEM is a microscope with powerful magnification that uses an electron beam accelerated by

300 kV, which can be transmitted through nanomaterials. The parts of a TEM are similar to those

of an SEM, but the principle of the TEM is different. In the TEM, the electron beam is transmitted

through the sample, which should be nanoscale in size and suspended on a Cu mesh grid. Images

are collected after the electrons have been transmitted through the sample. These images are

magnified after detection on a fluorescent screen. The TEM images and selected area electron

diffraction (SAED) patterns were obtained using a JEOL ARM-200F. The samples were first

dispersed in absolute ethanol by a sonication instrument for 10 min, and then one or two drops

were dropped on the Cu mesh grid. Finally, the TEM grid was dried in the air for one hour or

more.

3.3.3 X-ray powder diffraction

X-ray diffraction (XRD) is a technique that is used to characterize the crystal structure and phase

purity of samples. Mainly, the XRD diffractometer consists of three important parts: an X-ray

tube, a sample holder, and an X-ray detector. The X-ray tube generates monochromatic X-rays

with wavelength λ, which are focused on the sample surface. The interaction between the surface

and the monochromatic X-rays should produce constructive interference. The constructive

interference follows Bragg's Law (nλ = 2d sin θ), where θ is the diffraction angle and d is the

spacing of a set of lattice planes. The intensity of the diffracted rays is detected by the X-ray

detector, and the scanning angles vs. intensity provide crystallographic information. The data are

analysed by matching the data with data base that was collected before for many compounds.

In this work, XRD were carried out by X-ray diffraction (GBC MMA XRD) with the Cu Kα

wavelength of 1.542 Å and at 40 kV, 25 mA. The scan rate and range of angles were 1 deg/s and

10˚ to 60˚ respectively.

3.3.4 Ultraviolet-Visible-Near Infrared (UV-VIS-NIR) Spectrophotometer

The basic principle of this instrument is that light (covering a wide range from 100 to 1500 nm)

passes through a sample (which may be gas, liquid, or solid). The light intensity is analysed with a

monochromator and then detected by a photodetector after passing through the sample. The

absorptivity coefficients follow the Beer-Lambert law which is a proportional relationship between

the absorbance and the concentration of the sample. The solid powder was measured according to

principle of light beam interaction with a smooth sample, and the reflected light was collected in a

Page 56: P-block-based Ferroelectric-photocatalyst compounds

43

hemisphere shaped collector and then analysed.

Figure 3-1 Photograph of UV-Vix_NIR spectrometer and schematic diagram

showing the geometry in an integrating sphere.

In this work, a Shimadzu-3600 was used to measure diffuse reflectance spectra (DRS) of the

photocatalyst powder by adding an integrating sphere attachment (ISR-3100) to the instrument, as

shown in Figure 3.1, with BaSO4 providing a background between 200 nm and 800 nm. The dye

concentration in water after photodegradion was also measured by monitoring the peak absorption

of dye. The cuvette cell (4 ml in volume and optical path of 1 cm) was used to measure the dye

absorption in the range of 300-800 nm, and the reference cuvette cell was filled with pure water

for the background of the dye solution. In the case of phenol, it was used to measure absorption in

the range of 200-300 nm.

3.3.5 X-ray photoelectron spectroscopy (XPS)

Highly sensitive quantitative information on the elemental composition of a surface target was

detected by the XPS technique. The surface target is excited with X-rays, and then photon-electron

interaction occurs, with complete or partial photon energy transfer to core electrons. If this energy

is sufficient, then an electron is emitted from the sample surface with a given kinetic energy

(K.E.). This energy corresponds to the binding energy (B.E.) of the electron in an atom. The

electron detector measures the binding energy of the photoelectrons that are emitted from the

target under ultra-high vacuum (pressure, P < 10−9 millibar), and therefore, each peak position

reflecting a given binding energy of electrons corresponds to the electronic and oxidation state of

an element. The binding energy of the core electrons are affected by the electrons bonded with

surrounding atoms, and then a chemical shift occurs, which provides information on chemical

states. Besides this, the binding energy provides information about the band structure in

heterojunctions because the core electrons are affected by surrounding atoms which have different

electronegativity. In this work, XPS (PHI660) was performed using a monochromatic Mg Kα X-

ray source.

3.3.6 Atomic force microscopy (AFM) and piezoresponse force microscopy (PFM)

AFM is one of the most advanced instruments to measure the 3D topography of a sample with

Page 57: P-block-based Ferroelectric-photocatalyst compounds

44

ultra-high resolution. Nanoscale areas with 40-50 individual atoms can be studied by AFM to

determine their crystallographic properties, and it can be used for microscale observations, or even

larger than 100 µm, to show living cells. The basic principle is depends on two parts: a mechanical

part and an electrical part. A sharp probe maps the height of a surface sample with very close

contact to provide information on 3D topography. The movement of the sample is controlled by

piezoelectric materials. The vertical vibration of the probe on the surface sample is recorded by a

laser beam reflected from the probe surface, and then the laser beam is detected by a quadruple

photodiode. The digital signal recorded by the photodiode is analysed by controller and the

displayed on a computer [9]. The advantages of AFM are that samples do not need to be prepared

and can be observed under ambient atmosphere. AFM is also a low cost instrument compared with

SEM and TEM. In this work, Asylum Research AFMs were used to test thickness of nanosheets.

Piezoresponse force microscopy (PFM) was used for bismuth silicate nanosheets. The PFM can

study ferroelectric properties on the nanoscale, including the characteristics of hysteresis loops and

the switching of polarization. The basic principle is that voltage is applied on the probe tip and

results in the expansion of the sample when it is in contact with the sample surface. There are more

specific details reported in Chapter 5.

3.3.7 Polarization measurement

The ferroelectric materials have non-centrosymmetric crystal structures, which generate

polarization that can be reoriented by applying an external electric field. The polarization as a

function of the electric field was obtained using an Easy Check 300 (aixACCT Systems GmbH)

instrument equipped with a Trek 610E high voltage source, which is used for silver silicate.

3.3.8 Photoelectrochemical and electrochemical measurements

When a semiconductor is immersed in electrolyte in the dark, the charge density will be

redistributed in the interface region, and then the charges are transferred between the two materials

until an equilibrium state is reached. The Fermi level will be repositioned, and thus band bending

will be achieved. The region in the interface where this takes place is called the space-charge layer.

The relationship between the capacitance at the space charge layer and the external applied voltage

(as shown below in the Mott-Schoktty equation 3.1) can make it possible for us to obtain the flat

band potential and charge density, which are useful parameters to determine the conduction band

and valence band positions of the semiconductor [10].

Where

C is the capacitance of the space charge region,

kB is Boltzmann's constant

T is the temperature

q is the electronic charge

ԑ is the dielectric constant of the semiconductor,

1

𝐶2=

2

𝜀𝜀ₒ𝑁𝐷 𝐸 − 𝐸𝑓𝑏 −

𝑘𝐵𝑇

𝑞 3-1

Page 58: P-block-based Ferroelectric-photocatalyst compounds

45

ԑ0 is the permittivity of free space,

ND is the donor density (electron donor concentration for an n-type semi-conductor or hole

acceptor concentration for a p-type semi-conductor),

E is the applied potential,

EFB is the flatband potential.

A standard three-electrode system was adopted in this system, and the photocatalyst films,

platinum wire, Ag/AgCl, and 0.1 M Na2SO4 were used as the working electrode, counter electrode,

reference electrode, and electrolyte, respectively. The working electrode was prepared via the dip-

coating method on indium tin oxide (ITO) and had an area of 1 cm2 and frequency of 5 kHz. .The

Mott–Schottky curves were collected with a VSP-300 potentiostat.

In the photocurrent–time response system, a 300 W Xe lamp with a monochromator and a cut-off

filter (λ > 400 nm) was used as the light source. The photocurrent as a function of irradiation time

under visible light was collected with zero bias.

3.3.9 Photoluminescence (PL) spectra

One application of PL is to determine radiative and nonradiative recombination rates which is an

important investigation in photocatalysis. The basic principle is that a light beam is incident on a

sample and the sample is excited, so that it emits photons in a relaxation process, which equals the

energy between two electronic level. The charge separation was investigated by using

photoluminescence (PL) spectra, which were recorded on a Hitachi F-4500 fluorescence

spectrophotometer at room temperature and with an excitation wavelength of 320 nm.

3.3.10 Surface photovoltage (SPV)

SPV is an important technique to obtain band bending, recombination lifetimes, and interfacing

states of p-n junctions, as well as the minority carrier diffusion length of semiconductors and many

more characterizations. The SPV technique investigates the potential of a semiconductor surface

while excitation light is applied on surface. Usually, SPV is carried out by preparing a thin film of

the semiconductor material on conductive electrode and then applying light with a monochromatic

wavelength on the surface of the thin film. A Kelvin probe can usually scan the whole surface

without contact. The photovoltage can be investigated by detecting photoinduced charges carriers

that reach the semiconductor surface before and after it is illuminated with light equal to or higher

than the band gap [11].

3.4 Photocatalytic activity evaluation

The photocatalytic activity was evaluated by the decolouration of dyes and photodegradation of

phenol, which were used as model organic pollutants in the photocatalytic degradation experiment

at room temperature. A 300 W Xe lamp was used as the simulated light source for studying the

photocatalytic performance. In a typical process, 50 mL of RhB (10 mg l−1) (for phenol 20 mg l−1)

and 0.05 g of the as-prepared sample were added together in a 100 mL beaker with a diameter of 6

cm. The suspension was magnetically stirred in the dark for 30 min to reach the adsorption–

desorption equilibrium, and then the mixture was irradiated with the light source. At certain time

intervals, a 3 mL aliquot was taken and centrifuged to remove the photocatalyst. The

Page 59: P-block-based Ferroelectric-photocatalyst compounds

46

photodegradation activity was analysed using the UV-Vis spectrophotometer by recording the

variation of the peak absorption of dyes and phenol.

3.4.1 Determine reaction kinetics kt and quantum efficiency

To calculate the reaction kinetics kt of dyes and phenol degradation, the pseudofirst-order kinetic

model was derived, as shown in the equation below.

Kt =Ln C0

Ln Ct

Where k is removal rate constant, C0 is the initial dye concentration, and Ct is dye concentration at

time t after the start of the photocatalysis process.

The apparent quantum efficiency ( ) can be expressed as following equation:

𝜑 =No.of molecules undergo photodegradation

No.of photons accident inside reaction∗ 100

3.5 Programs: analysis

The programs were used in this work are listed below in Table 3.3 along with their purpose.

Table 3-3 Programs used in this work.

Program

name Ver. Purpose

VESTA 3.4.3 3D visualization of volumetric data, crystal

structure and determine the planes

Origin OriginPro Data analysis and graphing

Gatan

Digital

Micrograph

2.11.1404.0

Visualization and analysis of TEM and SAED

images, and determining the dspacing value and

exposed facetd of crystals

Match 2.3.2 Identify one or more phases of materials and

crystal structure from diffraction data.

NSS 3.3 Analyze the EDS spectrum to find the

composition of a sample

Casa ver.

2.3.16dev85 XPS analysis

3.6 Theoretical calculations

Theoretical calculations have become essential to support experimental results and directly find

properties of any solid. It has recently become widespread in the study of the chemical and

physical properties of materials for the interpretation and prediction of any solid containing

multiple electrons. The band structure, density of states, and polarization for many materials can

3-2

3-1

Page 60: P-block-based Ferroelectric-photocatalyst compounds

47

be calculated theoretically. It is thus possible to find better materials for batteries, solar cells,

photocatalysts, and many other applications. The electronic structure is achieved by solving

quantum mechanical equations for a many-electron system. Density functional theory (DFT) is a

computational method to solve many-electron systems. This calculation is not very accurate,

however, and sometimes not consistent with experimental results. To overcome this source of

error, many approximations for the exchange-correlation energy are used to improve the results.

Usually, the calculation is obtained using a supercomputer and appropriate software programs.

There are several separate programs that are used in calculations such as Vienna Ab-initio

Simulation Package (VASP).

3.7 References

1. Jansen, M. and H.H. Käs, Amorphous Silver Silicates, Storage Elements with Low Switching

Field Strength. Angewandte Chemie International Edition in English, 1980. 19(5): p. 386-387.

2. Beberwyck, B.J., Y. Surendranath, and A.P. Alivisatos, Cation Exchange: A Versatile Tool for

Nanomaterials Synthesis. The Journal of Physical Chemistry C, 2013. 117(39): p. 19759-19770.

3. Beberwyck, B.J. and A.P. Alivisatos, Ion Exchange Synthesis of III–V Nanocrystals. Journal of

the American Chemical Society, 2012. 134(49): p. 19977-19980.

4. Li, J., Q. Wu, and J. Wu, Synthesis of Nanoparticles via Solvothermal and Hydrothermal

Methods, in Handbook of Nanoparticles, M. Aliofkhazraei, Editor. 2015, Springer International

Publishing: Cham. p. 1-28. 5. Kimura, T., Molten Salt Synthesis of Ceramic Powders, in Advances in Ceramics - Synthesis and

Characterization, Processing and Specific Applications, C. Sikalidis, Editor. 2011, InTech:

Rijeka. p. Ch. 04.

6. Liu, X., N. Fechler, and M. Antonietti, Salt melt synthesis of ceramics, semiconductors and

carbon nanostructures. Chemical Society Reviews, 2013. 42(21): p. 8237-8265.

7. Md Mehebub, A., G. Sujoy Kumar, S. Ayesha, and M. Dipankar, Lead-free ZnSnO3 /MWCNTs-

based self-poled flexible hybrid nanogenerator for piezoelectric power generation.

Nanotechnology, 2015. 26(16): p. 165403.

8. Tang, Q.-Y., Y.-M. Kan, P.-L. Wang, Y.-G. Li, and G.-J. Zhang, Nd/V Co-Doped Bi4Ti3O12

Powder Prepared By Molten Salt Synthesis. Journal of the American Ceramic Society, 2007.

90(10): p. 3353-3356.

9. Eaton, P. and P. West, Atomic Force Microscopy. 2010: OUP Oxford. 10. Gelderman, K., L. Lee, and S.W. Donne, Flat-Band Potential of a Semiconductor: Using the

Mott-Schottky Equation. J. Chem. Educ., 2007. 84(4): p. 685.

11. Cavalcoli, D. and A. Cavallini, Surface photovoltage spectroscopy ‐ method and applications.

physica status solidi c, 2010. 7(5): p. 1293-1300.

Page 61: P-block-based Ferroelectric-photocatalyst compounds

48

Page 62: P-block-based Ferroelectric-photocatalyst compounds

49

Chapter 4

A ferroelectric photocatalyst Ag10Si4O13 with visible-light

photooxidation properties

4.1 Introduction

Photocatalysis is one of the most promising ways to solve the energy crisis and environmental

pollution issues by using solar energy [1-11]. A number of photocatalysts have been proposed and

prepared using materials design and band engineering techniques. Some have achieved excellent

photocatalytic performance in either water splitting or the elimination of organic pollution.

Nevertheless, most of the high-performance photocatalysts proposed so far are transition metal

oxide (TMO) based materials. Generally, TMOs have wide band gaps (> 3.0 eV) due to partially

filled d-orbitals [12-15] which limits the absorption of visible light. In addition, their valence

bands (VBs) and conduction bands (CBs) are less dispersive due to the isotropic d orbitals of

transition metals (TM). The photo-induced charge carriers are hence given a large effective mass,

which hinders charge separation and mobility, and results in low quantum conversion efficiency in

photocatalysis, especially in the visible light spectrum. Consequently, their visible-light

photocatalytic performances are not as high as desired. Therefore, the exploration and

development of new visible-light responsive photocatalysts with high quantum conversion

efficiency through materials design and band engineering have emerged as an urgent and

challenging task in the field of photocatalysis.

Recently, AgO-based compounds have been reported with excellent visible-light-driven

photocatalytic activities [16-24]. Because of their unique electronic configurations. Since the d

orbitals of Ag are fully filled, the p orbitals in these compounds play an important role in the

energy levels and dispersion of the VB and CB. Generally, p orbitals lift the top of the VB and/or

lower the bottom of the CB in AgO-based photocatalysts, and promote generation of photo-

induced electrons and holes under visible light. Moreover, highly dispersive VBs or CBs can be

formed due to anisotropic p or hybridized sp states. As a result, the effective mass of photo-

induced charge carriers is much smaller in AgO-based compounds, which facilitates charge

transfer in the photocatalytic process under visible-light irradiation [25, 26]. Nevertheless, the

narrow band gap and dispersive electronic structure also facilitate the recombination of electrons

and holes in the photocatalytic process and, in turn, would limit the photocatalytic activities of

these compounds. Therefore, AgO-based compounds with efficient charge separation in

photocatalysis are still desired, although their visible-light photocatalytic performance is much

higher than those of the TMOs.

Among these AgO-based compounds, an Ag-incorporated p-block family, silver silicate

(AgxSiyOz),[19, 20, 24] has captured our attention recently. Besides the unique d10 state of Ag,

Page 63: P-block-based Ferroelectric-photocatalyst compounds

50

both Si and O in AgxSiyOz have typical sp/p electronic configurations, which make the band gap,

and the VB and CB in AgxSiyOz very likely to conform to the optimum electronic structure for

visible-light-driven photocatalysts. In particular, an internal electric field is expected in AgxSiyOz

compounds, owing to the spontaneous electronic polarization induced by the distorted tetrahedral

unit (SiO4) arrangement [27-29]. This internal electric field can effectively promote spatial charge

separation, which addresses the electron-hole recombination issues that are found in the other

AgO-based compounds. Moreover, the strength of the internal electric field can be feasibly

modulated by the number of distorted arrangements of SiO4 tetrahedral, and enhances the

photocatalytic activity of silver silicate. As a consequence, AgxSiyOz is regarded as a promising

compound family for exploring new visible-light-driven photocatalysts because of this unique

crystal and electronic structures. Nevertheless, there have been very few investigations aimed at

the design and development of AgxSiyOz photocatalysts so far.

In this work, we report a new visible-light-driven silver silicate photocatalyst, Ag10Si4O13, which

demonstrates high photocatalytic activity towards the elimination of organics. It was found that the

excellent photocatalytic performance of Ag10Si4O13 originates from its anisotropic and highly

dispersive band structure due to its p-block electronic configuration. Moreover, Ag10Si4O13

exhibits a large ferroelectric polarization due to very large SiO4 chains compare to the other SiO4-

based ferroelectric. It induces an intrinsic internal electric field in Ag10Si4O13, which further

enhances the photocatalytic activity by effective separation of photo-induced charges. Our

experimental and theoretical results suggest that silver silicates are promising candidate materials

to explore visible-light-active photocatalysts through structural and electronic engineering.

4.2 Experimental section

Sample preparation. In a typical experimental process, AgNO3 and Na2SiO3 (99.9%, Sigma-

Aldrich Company) were mixed in a mortar in a molar ratio of 2:1 and then pressed into a disk 1

cm in diameter and 3 mm in thickness. The disk was heated in air at 400 ºC for 2 h. After that, the

product was washed three times with distilled water and ethanol to remove residual NaNO3 and

AgNO3. The sample was dried at 60 ºC for 12 h. For comparison, the photocatalytic activity of N

doped TiO2 was synthesized according to the original report [2].

4.2.1 Characterizations

The purity and crystallinity of the sample were investigated by X-ray diffraction (XRD, GBC,

MMA) using Cu Kα radiation with λ = 1.5418 Å. The sample morphology and elemental analysis

were investigated by field emission scanning electron microscopy (FESEM, JEOL-7500). The

composition of the samples was examined by energy dispersive spectroscopy (EDS) attached to

the FESEM. Transmission electron microscope (TEM) images, High-resolution TEM (HRTEM)

and Selected area electron diffraction (SAED) patterns were obtained using a JEOL ARM-200F.

X-ray photoelectron spectroscopy (XPS, PHI660) was performed using a monochromatic Mg Kα

X-ray source. An ultraviolet-visible spectrophotometer (UV-Vis, Shimadzu-3600) was used to

measure diffuse reflectance spectra (DRS) by adding an integrating sphere attachment to the

instrument, with BaSO4 providing a background between 200 nm and 800 nm. The polarization as

Page 64: P-block-based Ferroelectric-photocatalyst compounds

51

a function of electric field was obtained using an Easy Check 300 (aixACCT Systems GmbH)

equipped with a Trek 610E high voltage source. In the photocurrent-time response system, a 300

W Xe lamp with a monochromator and a cut-off filter (λ > 400 nm) was used as the light source.

The photocurrent as a function of irradiation time under visible light was collected by a

KEITHLEY 2400 source meter. The surface photovoltage (SPV) spectroscopy apparatus is

composed of a source of monochromatic light, a lock-in amplifier (SR830-DSP) with a light

chopper (SR540), and a photovoltaic cell. A 500 W Xe lamp (CHFXQ500W, Global Xenon Lamp

Power) and a grating monochromator (Omni-5007, no. 09010, Zolix) provide monochromatic

light. The construction of the photocurrent-time-response cell and the photovoltaic cell was in the

form of a sandwich-like structure of indium tin oxide (ITO)-sample-ITO. The Mott-Schottky

curves were collected with a PARSTAT-2273 Advanced Electrochemical System (Princeton

Applied Research). The Ag10Si4O13 film, Pt foil, SCE, and saturated KCl solution were used as

working, counter, and reference electrodes, and as the electrolyte, respectively. The working

electrode was prepared via the dip-coating method. The Mott-Schottky measurements were

monitored at a fixed frequency of 100 Hz with 10 mV amplitude at various potentials.

Photocatalytic activity measurements. Rhodamine B (RhB), methyl orange (MO), methylene blue

(MB), and phenol were used as target organics in photocatalytic degradation tests. All the dyes

and the phenol were purchased from Sigma-Aldrich. The photocatalytic activity was carried out at

room temperature using Ag10Si4O13 as photocatalyst. The light source used in the photocatalytic

measurements was a 300 W W-Xe lamp with a UV cut-off filter (λ > 420 nm). In a typical

process, 100 ml of RhB, MO, or MB solution (10 mg/L for all dyes) and 0.1 g Ag10Si4O13 were

added together in a 250 ml beaker. The suspensions were magnetically stirred in the dark for 30

min to reach adsorption-desorption equilibrium, and then the mixture was exposed to the light

source. The photodegradation of the dye solutions was analysed by UV-Vis spectrophotometer. To

stand in for real organic pollutants, phenol was used under the same conditions, but the

concentration and volume were 20 mg/L and 50ml respectively. The suspensions were

magnetically stirred in the dark for 120 min to reach adsorption-desorption equilibrium, a sample

taken every 20 min. For comparison, the photocatalytic activity of N doped TiO2 was measured

under the same conditions.

Theoretical calculations: The calculations were carried out using the Vienna Ab-initio Simulation

Package (VASP) based on density functional theory with the projector augmented wave (PAW)

pseudopotential method. We applied hybrid functional calculations to simulate ferroelectricity in

this work. The short-range exchange potential was calculated by mixing a fraction of nonlocal

Hartree–Fock exchange with the generalized gradient approximation (GGA) functional of Perdew,

Burke, and Ernzerhof (PBE). A conjugate-gradient algorithm was used to relax the ions into their

ground states, and the energies and the forces on each ion were converged within 1 ×10-5 eV/atom

and 0.01 eV/Å, respectively. The Koln-Sham orbitals were expanded by a plane wave basis set,

and an energy cut-off of 550 eV was used throughout. The Brillouin-zone integration was

performed by using the Gamma-centered Monkhorst-Pack scheme with 6 × 4 × 3 k-points.

4.3 Results and discussion

Page 65: P-block-based Ferroelectric-photocatalyst compounds

52

The XRD pattern of the as-prepared sample is shown in Figure 4-1. All the diffraction peaks

can be indexed to Ag10Si4O13 with a triclinic structure (space group P-1), according to the Joint

Committee on Powder Diffraction Standards (JCPDS) Card No. 01-071-1365. The sample

consists of phase-pure reddish powders without any observable impurities. In the triclinic

structure of Ag10Si4O13, as illustrated in the inset of Figure 4-1, four tetrahedral SiO4 units are

connected through a corner-shared oxygen atom, forming a distorted Si4O13 chain which is

coordinated with the Ag ions.

Figure 4-1 XRD pattern of the as-prepared Ag10Si4O13. The inset is the crystal

structure of Ag10Si4O13.

The FESEM images reveal that the as-prepared Ag10Si4O13 sample consists of micron-sized

irregular particles (Figure 4-2). As shown in Figure 4-3 (a) and (b), TEM image, HETEM and the SAED

pattern verifies excellent crystallinity of the Ag10Si4O13 sample. HRTEM results (Figure 4-3 (a)) clearly

show the lattice fringes of as-prepared Ag10Si4O13 particles with d-spacing of 0.197 and 0.279 nm,

which can be assigned to (-104) and (-411) directions of triclinic structure, respectively.

Page 66: P-block-based Ferroelectric-photocatalyst compounds

53

Figure 4-2 SEM image of the as-prepared Ag10Si4O13

Figure 4-3 Figure 4-3 (a) TEM image and (b) HETEM image and inset SAED pattern

of Ag10Si4O13.

Page 67: P-block-based Ferroelectric-photocatalyst compounds

54

Figure 4-4 Figure 4-3 (a) XPS result of the as-prepared Ag10Si4O13 (b) Ag 3d, (c) O

1s, (d) Si 2p. The spectra demonstrate that the main peaks correspond to Ag 3d5/2 and

Ag 3d3/2, O 1s, and Si 2p orbitals for the Ag10Si4O13.

The elemental composition and chemical states of Ag10Si4O13 were further analysed by

XPS, as shown in Figure 4-3 (a). The survey spectrum demonstrates that the main peaks

correspond to Ag 3d5/2 and Ag 3d3/2, O 1s, and Si 2p orbitals for the Ag10Si4O13 sample.

Detailed spectra providing the fine information for Ag, Si, and O are presented in Figure 4-

3 (b), (c) and (d) in which the peaks of Ag 3d5/2 and Ag 3d3/2 are located at 367 eV and 373

eV, whereas the peaks of Si 2p and O 1s are located at 100.8 eV and 530.6 eV, respectively.

The valences of Ag and Si are thus identified as +1 and +4, respectively.

The energy dispersive spectroscopy (EDS) spectrum shows (Figure 4-5) that the atomic

percentages of Ag, Si, and O in the sample are 26%, 12%, and 42%, respectively as shown

in inset Figure 4-5, which further confirms the composition of phase-pure Ag10Si4O13.

Page 68: P-block-based Ferroelectric-photocatalyst compounds

55

Figure 4-5 Figure 4-5 EDS spectrum of elements and inset table of elements

content percentage corresponding to the Ag10Si4O13.

Figure 4-6 shows the diffuse reflectance spectrum of Ag10Si4O13 collected by ultraviolet-

visible (UV-Vis) spectroscopy. The absorption edge is located at about 700 nm. The band

gap (Eg) of Ag10Si4O13 was then estimated to be 1.72 eV according to the Tauc formula

[30], as is shown in the upper right inset of Figure 4-6. It indicates that Ag10Si4O13 can

absorb almost all the visible-light spectrum. The narrow band gap of the as-prepared

Ag10Si4O13 sample is believed to have arisen from the hybridization of different orbitals in

this Ag-incorporated p-block compound. Thus, density functional theory (DFT)

calculations were carried out in order to reveal the electronic structure of Ag10Si4O13. The

calculated band structure is shown in Figure 4-7. The band gap is calculated as 1.71 eV,

which agrees well with our experimental results based on the photo-absorption edge.

The high symmetry points of the CB and VB are located at the Γ and Ζ points,

respectively, giving Ag10Si4O13 the features of a typical indirect band-gap semiconductor. It

is well known that the excited electrons in an indirect-band-gap semiconductor have a

longer lifetime. The separation and migration of photo-induced charge carriers are thus

expected to be more efficient. Moreover, a typical signature band structure of the p-block

configuration is revealed by our DFT results, in which the CB is highly dispersive whereas

the VB shows much less dispersion in Ag10Si4O13.

Page 69: P-block-based Ferroelectric-photocatalyst compounds

56

Figure 4-6 UV-vis diffuse reflectance spectra of Ag10Si4O13. Inset at the top right of (d) is the derivation of the band gap value; Inset at the bottom left of figure is a

photograph of the photocatalyst.

Figure 4-7 DFT calculated electronic structure of Ag10Si4O13, in which a dispersive

CB and flat VB can be observed

As shown in Figure 4-8, the CB of Ag10Si4O13 mainly consists of Ag 5s and 5p orbitals.

The large amount of hybridization between Ag s and Ag p occurs due to short Ag-Ag

bonds. This indicates that the CB is constructed without the “contamination” of less

Page 70: P-block-based Ferroelectric-photocatalyst compounds

57

dispersive d states. As a result, the photo-excited electrons have a small effective mass,

which enables photo-excited electrons to easily move to surface active sites due to their

high mobility. In contrast, the VB of Ag10Si4O13 is mainly contributed by Ag 4d orbitals.

This leads to a flatter band (as compared to the CB) and thus generates “heavy” holes in the

photo-excitation process. This electronic configuration suggests high quantum conversion

efficiency for Ag10Si4O13 under visible light.

Figure 4-8 Partial density of states calculated by DFT.

Page 71: P-block-based Ferroelectric-photocatalyst compounds

58

Figure 4-9 4-9 Mott–Schottky plot of pure Ag10Si4O13. The flat band potential is

determined to be about 0.523 V. The inset is a schematic diagram of the redox

potential of Ag10Si4O13, corresponding to the Mott–Schottky fitting result

schematic of band-edge potentials of Ag10Si4O13.

It is well known that the energy positions of the CB and VB are the primary factors that

determine whether the semiconductor can be used in photocatalytic reduction or oxidation. We

therefore carried out Mott-Schottky measurements of Ag10Si4O13, as shown in Figure 4-9. The

relative position of the CB edge can be calculated from the empirical equation [31]: Ec = χ − 0.5

Eg+ E0, where E0 is a scale factor relating the reference electrode redox level to the absolute

vacuum scale (E0 = −4.5 eV for normal hydrogen electrode (NHE)), Eg is the band gap, and χ is

the absolute electronegativity of the semiconductor. The calculated Ec from this equation is

empirical and theoretical. Therefore, electrochemical flat-band potential measurements were

adopted. The flat-band potential values are obtained by the Mott-Schottky equation [32, 33], as

shown in Figure 4-9. The flat potential is calculated to be 0.523 V versus saturated calomel

electrode (SCE), which is equivalent to 0.729 V (NHE vs. Ag/AgCl = 0.206 V). It is well

accepted that the CB of n-type semiconductors is about -0.1 V higher than the flat potentials,

depending on the electron effective mass and carrier concentration. With an applied voltage

difference of 0.05 eV between the conduction band and the flat potential, the bottom of the CB is

derived as 0.679 eV. The inset in Fig. 2c shows a schematic diagram of the redox potential of

Ag10Si4O13, corresponding to the Mott-Schottky fitting result. The CB and VB edges (inset in

Figure 4-9) of Ag10Si4O13 are +0.679 eV and +2.4 eV vs. NHE, respectively. The lower potential

of the VB with respect to the O2/H2O redox potential indicates that Ag10Si4O13 possesses strong

photo-oxidative capabilities and can be used as an oxidative photocatalyst.

Page 72: P-block-based Ferroelectric-photocatalyst compounds

59

Figure 4-10 UV-Visible absorbance spectra for the

photodegradation of (a) MO, (b) RhB and (c) MB under

visible light over Ag10Si4O13 recorded after different

degradation times. The insets shows the color changes

of the (a) MO, (b) RhB and (c) MB solutions

corresponding to the five degradation times from 0 min

to 40 min

In view of the possible strong photo-oxidative capability, the photocatalytic activity of

Ag10Si4O13 was evaluated by the photodegradation of organic compounds, including Rhodamine

B (RhB), Methyl orange (MO), methylene blue (MB), and phenol, under visible light irradiation.

The results are shown in Figure 4-10 (a), (b) and (c).

Page 73: P-block-based Ferroelectric-photocatalyst compounds

60

All the photocatalytic degradation measurements were carried out after the dark reaction, in

which the equilibrium adsorption states were reached. The Ag10Si4O13 sample exhibits excellent

visible-light photocatalytic degradation activity, which is better than that of N doped TiO2 and

the other AgO-based reference samples.

For example, under visible-light irradiation, the absorption peak of MB at 664 nm decreased

rapidly with increasing irradiation time and eventually disappeared after 40 min, suggesting

excellent photocatalytic performance of the Ag10Si4O13 (Figure 4-10 (c)). This demonstrates that

the MB can be completely decolourisation by Ag10Si4O13 under visible light. The inset in Figure

4-10 shows the visible changes in the color concentration of the MB, RhB and MO solution

during the photo-oxidation process. The excellent photocatalytic degradation activities of

Ag10Si4O13 were also demonstrated with the other dyes, which are shown in Figure 4-10 (a) and

(b). By monitoring the signature absorption peaks of RhB, MB, and MO at 556 nm, 664 nm, and

470 nm, respectively, plots of the visible-light degradation ratio versus reaction time were

obtained and are shown in Figure 4-11. We note that the complete decolourisation of MB and

RhB, and the 80% degradation of MO can be achieved after 30 min, 40 min, and 40 min,

respectively. The photocatalytic degradation reaction of the samples followed the pseudo-first-

order linear relationship with reaction rate constants (k) as shown in the inset of Figure 4-11.

Figure 4-11 photodegradation rate of organic dyes over Ag10Si4O13 under visible light.

Page 74: P-block-based Ferroelectric-photocatalyst compounds

61

Figure 4-12 Cycling runs in the photodegradation of RhB over Ag10Si4O13.

Figure 4-13 Comparison of the photodegradation percentages of Ag10Si4O13 and N

doped TiO2 (N–TiO2) under visible light for different organic dyes.

The photocatalytic stability of Ag10Si4O13 in photodegradation was also studied by a

photocatalytic cycling test, as shown in Figure 4-12. Ag10Si4O13 shows stable photocatalytic

Page 75: P-block-based Ferroelectric-photocatalyst compounds

62

performance for eight photocatalytic degradation cycles. The decreased photocatalytic activity in

the first several rounds is due to initial instability of Ag10Si4O13, which is general for the other

Ag-based photocatalysts [34, 35]. It is believed that a small amount of Ag nanoparticles form on

the surface of Ag10Si4O13 at the very beginning photocatalytic cycles. These Ag nanoparticles

then function as electron trapping centres to suppress further photocorrosion of the samples,

which is evident by stable photocatalytic performance after third run. It should be noted that our

Ag10Si4O13 sample shows much higher photo-oxidation activity than the N doped TiO2 reference

sample under visible-light, as shown in Figure 4-13, which verifies its excellent visible-light

photocatalytic activity towards degradation of organics. We also carried out photocatalytic

degradation of colorless phenol over our Ag10Si4O13 sample under visible light irradiation. As

shown in Figure 4-14 and Figure 4-15, it is found that the phenol can be completely degraded in

80 min. The inset in Figure 4-14 shows the plots of the rate constants (k) versus reaction time.

The above results confirm that the photocatalytic dye degradation activity of Ag10Si4O13 cannot

only be attributed to the dye-sensitization effect because the colorless phenol only absorbs light

with wavelengths of 260-280 nm. The k of MB, MO, RhB and phenol are 6, 2.68, 5.7, 2 s-1

respectively, as shown in Table (1). The apparent quantum efficiency are 0.78, 0.83, 1.23, 4.4

respectively [36].

Figure 4-14 Photodegradation rate of phenol over Ag10Si4O13 under visible light. Data points represent an average of duplicate measurements, and error bars

represent one standard deviation; the inset shows the kinetic study of the

photocatalytic degradation process of phenol under visible light over Ag10Si4O13.

Page 76: P-block-based Ferroelectric-photocatalyst compounds

63

Figure 4-15 High performance liquid chromatography (HPLC) spectra of phenol

under visible light over Ag10Si4O13

Table 4-1 the φ value for different organic pollutants

The mechanism behind the excellent visible-light photocatalytic activity of our Ag10Si4O13

sample was investigated. It is well known that the photocatalytic activity is dominated by several

chemical dynamic processes in photocatalysis, including charge excitation, charge separation,

and charge transfer. In our photocatalytic degradation measurements, Ag10Si4O13 shows superior

photocatalytic activity relative to N doped TiO2 and the other Ag-based p-block photocatalysts

(Figure 4-16). This indicates that one or several key steps involved in the photocatalytic

dynamics under visible light would be promoted by the electronic structure of Ag10Si4O13.

Page 77: P-block-based Ferroelectric-photocatalyst compounds

64

Figure 4-16 Photodegradation of phenol within 20 min over different Ag-incorporated p-block

photocatalysts. References: Ag2CO3[37], Ag3PO4 [38], Ag7Si2O7 [24], Ag10Si4O13 our work.

We carried out photocurrent response measurements of Ag10Si4O13 under visible light

irradiation, as shown in Figure 4-17. It can be seen that the photocurrent generated in Ag10Si4O13

under visible light is 2 × 10-1 mA/cm2 at the beginning of irradiation, which is higher than that (1

× 10-1 mA/cm2) of a typical visible-light-driven AgO-based photocatalyst Ag3PO4 [39].The ON-

OFF cycles measurement of photocurrent demonstrates that Ag10Si4O13 possesses high quantum

conversion efficiency, and there is a low recombination rate of photo-induced charge carriers

under visible-light illumination. It should note that gradually decrease of photocurrent at first

three cycles is due to photocorrosion of sample as a result of formation of Ag nanoparticles. We

went on to collect surface photovoltage spectra (Figure 4-18) which confirms the high separation

rate of photo-induced charge carriers in Ag10Si4O13, especially in the spectral range from 300 to

600 nm. The efficient separation of photo-induced charges is most likely the key factor

contributing to the excellent photocatalytic activity of Ag10Si4O13, as compared to the other p-

block photocatalysts.

Page 78: P-block-based Ferroelectric-photocatalyst compounds

65

Figure 4-17 Photocurrent density response with the light on/off over Ag10Si4O13.

.

Figure 4-18 Surface photovoltage spectrum of Ag10Si4O13, which shows its

highest response in the visible-light range.

In fact, the separation of electron-hole pairs in Ag10Si4O13 would not be expected to be so

effective if one only considered its electronic structure. The reasons for the greater efficiency are

as follows: Firstly, the charge recombination process also takes advantage of the dispersive band

structure in Ag10Si4O13. The chances for separation and recombination would be traded off in the

Page 79: P-block-based Ferroelectric-photocatalyst compounds

66

photocatalytic process. Secondly, our DFT results reveal that photo-induced holes in Ag10Si4O13

possess a large effective mass due to the relatively flat VB. This would be expected to limit the

movement of photo-induced holes to surface active sites of Ag10Si4O13, and thereby, limit the

photo-oxidative activity of Ag10Si4O13. Nevertheless, this is obviously contradictory to our

results on the photocatalytic degradation of dyes and phenol in Figure 4-11. Therefore, it is

suspected that the effective separation of photo-induced charge carriers in Ag10Si4O13 is caused

by another mechanism in addition to the p-block electronic configuration.

In AgxSiyOz compounds, an internal electric field may exist due to the distorted SiO4

tetrahedral chains that are formed in the crystal structure. This internal electric field can

overcome charge recombination and promote spatial separation of electrons and holes.

Interestingly, Ag10Si4O13 has the longest distorted SiO4 chains in the AgxSiyOz family. They most

likely induce an internal field due to spontaneous polarization. In order to verify this, a

ferroelectric polarization measurement was conducted. As shown in Figure 4-19, we observed the

ferroelectric polarization-electric field (P-E) hysteresis loops of Ag10Si4O13, which were obtained

through a pulsed positive-up-negative-down (PUND) polarization measurement. A typical

unsaturated ferroelectric hysteresis loop with a remanent polarization (Pr) of 0.05 µC/cm2 was

obtained in electric field of 100 V/cm, as shown in Figure 4-17. This result verifies the existence

of an internal electric field in Ag10Si4O13 due to an intrinsic electric polarization. Corresponding

to the DFT-simulated three-dimensional (3D) charge density distribution (Figure 4-20), the

spontaneous electric polarization in Ag10Si4O13 is attributed to misaligned positive and negative

charge centers in the Ag10Si4O13 unit cells, in which O is the charge acceptor and Si is the main

charge donor, with little charge decrease for Ag in the crystal. This result suggests that the

spontaneous electric polarization in Ag10Si4O13 is attributable to misaligned positive and

negative charge centers, which are mainly induced by the asymmetrical Si-O tetrahedral chains in

Ag10Si4O13 unit cells. Given this structure, the calculated spontaneous polarization is about 64.78

µC/cm2, and the polarization direction is indicated in Figure 4-21.

We also find that the O1-O1 bonds in the chains are twisted with angles of 143º and 166.4º

(Figure 4-22) accompanied by a coherent displacement of the Ag ions, as well as a distortion of

the O3. The SiO4 tetrahedra show a spiral distortion with respect to each other on the SiO4

chains. The spontaneous polarization along the c direction is induced by the relative displacement

of these anionic SiO4 tetrahedral units and cationic Ag-O layers. Concurrently, the Ag ions are

slightly shifted to a direction normal to the O3, thereby also creating electronic polarization along

the a direction. The internal electric field is believed to be the main driving force for charge

separation in Ag10Si4O13. Our investigation has demonstrated that the excellent visible-light

photocatalytic activity of Ag10Si4O13 is attributable to its p-block electronic structure as well as

structural distortion, which provides a new strategy to design and develop visible-light-driven

photocatalysts to eliminate pollution.

Page 80: P-block-based Ferroelectric-photocatalyst compounds

67

Figure 4-19 P–E hysteresis loop of Ag10Si4O13 at different applied

voltages and frequency 1 kHz.

Figure 4-20 Charge density map; the isosurfaces are 0.01e / Å3 of

Ag10Si4O13.

Page 81: P-block-based Ferroelectric-photocatalyst compounds

68

Figure 4-21 Crystal structure of Ag10Si4O13 to show polarization direction

Figure 4-22 Three-dimensional chain composed of four SiO4 tetrahedra.

Note that the bonds twist with angles of 166.4° and 143º.

4.4 Conclusions

The visible-light-driven photocatalyst Ag10Si4O13 was synthesized by the ion-exchange method.

It was found that Ag10Si4O13 possesses excellent photocatalytic performance towards degradation

of dyes and phenol over the entire visible light spectrum. The photocatalytic activity of

Ag10Si4O13 is attributed to high quantum conversion efficiency because of its narrow band gap,

Page 82: P-block-based Ferroelectric-photocatalyst compounds

69

dispersive band structure, and internal electric field. This unique electronic structure and internal

electric field originate from the p-block electronic configuration and distorted SiO4 chains in

Ag10Si4O13.

4.5 References

1. Fujishima, A. and K. Honda, Electrochemical Photolysis of Water at a Semiconductor Electrode.

Nature, 1972. 238(5358): p. 37-38.

2. Vaiano, V., O. Sacco, D. Sannino, and P. Ciambelli, Photocatalytic removal of spiramycin from

wastewater under visible light with N-doped TiO2 photocatalysts. Chem. Eng. J., 2015. 261(0): p.

3-8.

3. Sacco, O., V. Vaiano, C. Han, D. Sannino, and D.D. Dionysiou, Photocatalytic removal of

atrazine using N-doped TiO2 supported on phosphors. Appl. Catal., B, 2015. 164(0): p. 462-474.

4. Takeuchi, M., H. Yamashita, M. Matsuoka, M. Anpo, T. Hirao, N. Itoh, and N. Iwamoto,

Photocatalytic decomposition of NO on titanium oxide thin film photocatalysts prepared by an ionized cluster beam technique. Catal. Lett., 2000. 66(3): p. 185-187.

5. Klosek, S. and D. Raftery, Visible Light Driven V-Doped TiO2 Photocatalyst and Its

Photooxidation of Ethanol. J. Phys. Chem. B, 2001. 105(14): p. 2815-2819.

6. Khan, M., S.R. Gul, J. Li, W. Cao, and A.G. Mamalis, Preparation, characterization and visible

light photocatalytic activity of silver, nitrogen co-doped TiO2 photocatalyst. Mater. Res. Express,

2015. 2(6): p. 066201.

7. Wang, Q., M. Chen, N. Zhu, X. Shi, H. Jin, Y. Zhang, and Y. Cong, Preparation of AgI

sensitized amorphous TiO2 as novel high-performance photocatalyst for environmental

applications. J. Colloid Interface Sci., 2015. 448(0): p. 407-416.

8. Cho, K.M., K.H. Kim, H.O. Choi, and H.-T. Jung, A highly photoactive, visible-light-driven

graphene/2D mesoporous TiO2 photocatalyst. Green Chem., 2015. 17(7): p. 3972-3978.

9. Aslam, M., I.M.I. Ismail, S. Chandrasekaran, and A. Hameed, Morphology controlled bulk synthesis of disc-shaped WO3 powder and evaluation of its photocatalytic activity for the

degradation of phenols. J. Hazard. Mater., 2014. 276: p. 120-128.

10. Lang, D., Q. Xiang, G. Qiu, X. Feng, and F. Liu, Effects of crystalline phase and morphology on

the visible light photocatalytic H2-production activity of CdS nanocrystals. Dalton Transactions,

2014. 43(19): p. 7245-7253.

11. Shang, J., W. Hao, X. Lv, T. Wang, X. Wang, Y. Du, S. Dou, T. Xie, D. Wang, and J. Wang,

Bismuth Oxybromide with Reasonable Photocatalytic Reduction Activity under Visible Light.

ACS Catal., 2014. 4(3): p. 954-961.

12. Janáky, C., K. Rajeshwar, N.R. de Tacconi, W. Chanmanee, and M.N. Huda, Tungsten-based

oxide semiconductors for solar hydrogen generation. Catal. Today, 2013. 199: p. 53-64.

13. Meng, F., J. Li, Z. Hong, M. Zhi, A. Sakla, C. Xiang, and N. Wu, Photocatalytic generation of hydrogen with visible-light nitrogen-doped lanthanum titanium oxides. Catal. Today, 2013. 199:

p. 48-52.

14. Shu, Z., X. Jiao, and D. Chen, Hydrothermal synthesis and selective photocatalytic properties of

tetragonal star-like ZrO2 nanostructures. CrystEngComm, 2013. 15(21): p. 4288-4294.

15. Yan, H., X. Wang, M. Yao, and X. Yao, Band structure design of semiconductors for enhanced

photocatalytic activity: The case of TiO2. Prog. Nat. Sci., 2013. 23(4): p. 402-407.

16. Xu, C., Y. Liu, B. Huang, H. Li, X. Qin, X. Zhang, and Y. Dai, Preparation, characterization,

and photocatalytic properties of silver carbonate. Appl. Surf. Sci., 2011. 257(20): p. 8732-8736.

17. Ouyang, S., N. Kikugawa, Z. Zou, and J. Ye, Effective decolorizations and mineralizations of

organic dyes over a silver germanium oxide photocatalyst under indoor-illumination irradiation.

Appl. Catal., A, 2009. 366(2): p. 309-314.

18. Maruyama, Y., H. Irie, and K. Hashimoto, Visible Light Sensitive Photocatalyst, Delafossite Structured α-AgGaO2. J. Phys. Chem. B, 2006. 110(46): p. 23274-23278.

19. Kim, T.-G., D.-H. Yeon, T. Kim, J. Lee, and S.-J. Im, Silver silicates with three-dimensional d10-

d10 interactions as visible light active photocatalysts for water oxidation. Appl. Phys. Lett.,

2013. 103(4): p. 043904.

20. Zhu, X., Z. Wang, B. Huang, W. Wei, Y. Dai, X. Zhang, and X. Qin, Synthesis of

Ag9(SiO4)2NO3 through a reactive flux method and its visible-light photocatalytic

performances. APL Mater., 2015. 3(10): p. 104413.

Page 83: P-block-based Ferroelectric-photocatalyst compounds

70

21. Zhao, J., Z. Ji, C. Xu, X. Shen, L. Ma, and X. Liu, Hydrothermal syntheses of silver phosphate

nanostructures and their photocatalytic performance for organic pollutant degradation. Cryst.

Res. Technol., 2014. 49(12): p. 975-981.

22. Tang, J., D. Li, Z. Feng, Z. Tan, and B. Ou, A novel AgIO4 semiconductor with ultrahigh activity

in photodegradation of organic dyes: insights into the photosensitization mechanism. RSC

Advances, 2014. 4(5): p. 2151-2154.

23. Yu, K., S. Yang, C. Liu, H. Chen, H. Li, C. Sun, and S.A. Boyd, Degradation of Organic Dyes

via Bismuth Silver Oxide Initiated Direct Oxidation Coupled with Sodium Bismuthate Based

Visible Light Photocatalysis. Environ. Sci. Technol., 2012. 46(13): p. 7318-7326.

24. Lou, Z., B. Huang, Z. Wang, X. Ma, R. Zhang, X. Zhang, X. Qin, Y. Dai, and M.-H. Whangbo, Ag6Si2O7: a Silicate Photocatalyst for the Visible Region. Chem. Mater., 2014. 26(13): p. 3873-

3875.

25. Feng, H., Z. Xu, L. Wang, Y. Yu, D. Mitchell, D. Cui, X. Xu, J. Shi, T. Sannomiya, Y. Du, W.

Hao, and S.X. Dou, Modulation of Photocatalytic Properties by Strain in 2D BiOBr Nanosheets.

ACS Appl. Mater. Interfaces, 2015. 7(50): p. 27592-27596.

26. Wang, L., J. Shang, W. Hao, S. Jiang, S. Huang, T. Wang, Z. Sun, Y. Du, S. Dou, T. Xie, D.

Wang, and J. Wang, A dye-sensitized visible light photocatalyst-Bi24O31Cl10. Sci. Rep., 2014.

4: p. 7384.

27. Taniguchi, H., A. Kuwabara, J. Kim, Y. Kim, H. Moriwake, S. Kim, T. Hoshiyama, T. Koyama,

S. Mori, M. Takata, H. Hosono, Y. Inaguma, and M. Itoh, Ferroelectricity Driven by Twisting of

Silicate Tetrahedral Chains. Angew. Chem., Int. Ed., 2013. 52(31): p. 8088-8092. 28. Seol, D., H. Taniguchi, J.-Y. Hwang, M. Itoh, H. Shin, S.W. Kim, and Y. Kim, Strong

anisotropy of ferroelectricity in lead-free bismuth silicate. Nanoscale, 2015. 7(27): p. 11561-

11565.

29. Akira SAHASHI, T.H., Hiroaki TAKEDA and Takaaki TSURUMI, Fabrication of ferroelectric

silicate KNbSi2O7 single crystal. J. Ceram. Soc. Jpn., 2014. 122(1426): p. P6-1-P6-4.

30. Tauc, J., R. Grigorovici, and A. Vancu, Optical Properties and Electronic Structure of

Amorphous Germanium. Phys. Status Solidi B, 1966. 15(2): p. 627-637.

31. Butler, M.A. and D.S. Ginley, Prediction of Flatband Potentials at Semiconductor‐Electrolyte

Interfaces from Atomic Electronegativities. Journal of The Electrochemical Society, 1978.

125(2): p. 228-232.

32. Gelderman, K., L. Lee, and S.W. Donne, Flat-Band Potential of a Semiconductor: Using the Mott-Schottky Equation. J. Chem. Educ., 2007. 84(4): p. 685.

33. Cardon, F. and W.P. Gomes, On the determination of the flat-band potential of a semiconductor

in contact with a metal or an electrolyte from the Mott-Schottky plot. J. Phys. D: Appl. Phys.,

1978. 11(4): p. L63.

34. Xu, D., B. Cheng, J. Zhang, W. Wang, J. Yu, and W. Ho, Photocatalytic activity of Ag2MO4 (M

= Cr, Mo, W) photocatalysts. Journal of Materials Chemistry A, 2015. 3(40): p. 20153-20166.

35. Yi, Z., J. Ye, N. Kikugawa, T. Kako, S. Ouyang, H. Stuart-Williams, H. Yang, J. Cao, W. Luo,

Z. Li, Y. Liu, and R.L. Withers, An orthophosphate semiconductor with photooxidation

properties under visible-light irradiation. Nature Materials, 2010. 9: p. 559.

36. Buriak, J.M., P.V. Kamat, and K.S. Schanze, Best Practices for Reporting on Heterogeneous

Photocatalysis. ACS Applied Materials & Interfaces, 2014. 6(15): p. 11815-11816.

37. Yu, C., G. Li, S. Kumar, K. Yang, and R. Jin, Phase Transformation Synthesis of Novel Ag2O/Ag2CO3 Heterostructures with High Visible Light Efficiency in Photocatalytic

Degradation of Pollutants. Advanced Materials, 2014. 26(6): p. 892-898.

38. Jiang, B., Y. Wang, J.-Q. Wang, C. Tian, W. Li, Q. Feng, Q. Pan, and H. Fu, In Situ Fabrication

of Ag/Ag3PO4/Graphene Triple Heterostructure Visible-Light Photocatalyst through Graphene-

Assisted Reduction Strategy. ChemCatChem, 2013. 5(6): p. 1359-1367.

39. Liu, Y., L. Fang, H. Lu, Y. Li, C. Hu, and H. Yu, One-pot pyridine-assisted synthesis of visible-

light-driven photocatalyst Ag/Ag3PO4. Applied Catalysis B: Environmental, 2012. 115–116: p.

245-252.

Page 84: P-block-based Ferroelectric-photocatalyst compounds

71

Chapter 5

Enhancement of charge separation in ferroelectric heterogeneous

photocatalyst Bi4(SiO4)3/Bi2SiO5 nanostructures

5.1 Introduction

Among several strategies for promoting electron-hole separation, building an electric field

internally or band bending in the interface have been proved to be efficient ways to accelerate the

electron-hole separation and drive the carriers to the interface for the water splitting reaction or for

organic pollutant degradation [1-4]. Recently, the spontaneous polarization in a typical

ferroelectric semiconductor has been reported to enhance separation of photogenerated charge

carriers at the different facets of the crystal [5-7]. The generation and control of polarization from

ferroelectric domains are, therefore, expected to represent a new strategy for designing novel

photocatalysts, which may induce electric fields due to energy band bending of similar magnitude

to that of p-n junctions.

The silicate-based compounds, as a new ferroelectric materials family, have shown huge

potential for photocatalysis due to their high polarization and internal electric field induced by the

distorted tetrahedral unit (SiO4) arrangement [8-10]. In our previous work, we showed that the

silicate compound Ag10Si4O13 with silicate chains featuring p-orbital electronic configurations was

successfully prepared [11]. The p-block element-based dispersive band structures guaranteed high

mobility of the photoinduced carriers [12-15] and the strong internal electric field caused by the

distortion of the long silicate chains further accelerated the separation of these electron-hole

carriers, resulting in a highly effective photocatalytic performance. As another typical silicate

compound featuring p-orbital electronic configurations, Bi2SiO5 was expected to be a superior

photocatalyst because it also has highly dispersive band structures and highly ferroelectric

polarization. Nevertheless, Bi2SiO5 has been reported to show unsatisfactory photocatalytic

activity towards inorganic degradation [16-20]. Similar to other layered Aurivillius-type oxides,

Bi2SiO5 is crystallized in layered structures composed of [Bi2O2]2+ slabs interleaved with anion

layers and tends to expose a crystal plane perpendicular to the c-axis (parallel to the [Bi2O2]2+

layer) during the growing process for crystals or nanocrystals [21] . Unfortunately, the main

direction of the polarization and internal electric field is along the c-axis for Bi2SiO5 [22, 23]. This

means that, for most Bi2SiO5 nanostructures, especially Bi2SiO5 nanosheets, the photoinduced

electron-hole pairs could separate under the polarization but would have to travel long distances to

reach the active edge sites because of the large length to height ratio of the two-dimensional

crystal structure. The separated charges are likely to be trapped by defects along the diffusion

path, and the spatial charge separation driven by the polarization could not significantly promote

the photocatalytic reaction occurring on the surface of Bi2SiO5.

In this work, we overcome and utilize the intrinsic disadvantage of the polarization direction in

bismuth silicate nanosheets for photocatalysis, by constructing unique zero-dimensional – two-

Page 85: P-block-based Ferroelectric-photocatalyst compounds

72

dimensional (0D-2D) Bi4(SiO4)3/Bi2SiO5 heterogeneous nanostructures. The heterostructures were

successfully synthesized by in-situ growth of Bi4(SiO4)3 nanodots on the Bi2SiO5 nanosheets via a

one-pot hydrothermal process. It is the connection between these two bismuth silicate compounds,

and thus, the resultant band bending at the interface, which drives the photogenerated charge

carriers in these two semiconductors to the relevant sides [24]. Moreover, a photocatalyst

heterostructure build-in the same elements but different phases has been shown significantly

enhancement photocatalytic activity [25, 26]. Integrating the internal polarization direction with

the heterostructure-induced charge-separation direction, the photoinduced charge carriers in this

heterogeneous structure are effectively separated and are driven to the surface rather than to the

edge sites, significantly enhancing the photocatalytic performance. The p-orbital electronic

configurations that are characteristic of both bismuth silicates here guarantee the high mobility of

the carriers, which also contributes to the charge separation and photocatalytic performance.

5.2 Experimental

5.2.1 Synthesis of bismuth silicate

All chemicals were analytical grade and used as received from Sigma-Aldrich without further

purification. The Bi4(SiO4)3/Bi2SiO5 nanosheet (BSO-HNS) 0D-2D heterostructures were

synthesized by the hydrothermal method. 0.1 g of cetyltrimethylammonium bromide (CTAB) and

0.1 g of polyvinylpyrrolidone (PVP) (30 K) were pre-dissolved in 5 ml distilled water and then

mixed with 0.05 M Bi(NO3)3∙5H2O solution. 0.5 mmol Na2SiO5 was added to the above mixture

under magnetic stirring. The pH value of the solution was then adjusted to 10 by adding NaOH.

After stirring for 10 min, the mixture was transferred into a stainless-steel Teflon-lined autoclave

with 100 mL capacity. The autoclave was sealed and heated at 200 °C for 6 h and then cooled to

room temperature. The collected products were centrifuged, rinsed with distilled water and

ethanol, respectively, and then dried at 60 °C for 5 h.

For comparison, our samples with pure phase Bi2SiO5 (BSO1) and Bi4(SiO4)3 (BSO2) were

synthesized under the same conditions. 1 mmol Bi(NO3)3∙5H2O was added into 20 mL mannitol

solution (0.1 M) under sonication and stirred for 15 min until the solution became clear. 0.5 mmol

Na2SiO5 was then added to obtain BSO1 (1 mmol for the synthesis of BSO2). The pH values of

the solutions were then adjusted to 10 by adding NaOH. The mixture was transferred into a

stainless-steel Teflon-lined autoclave with 100 mL capacity. The autoclave was sealed and heated

at 180 °C for 12 h (24 h for BSO2), and then cooled to room temperature naturally. The collected

products were centrifuged, rinsed with distilled water and ethanol, respectively, and then dried at

60 °C for 5 h.

5.2.2 Characterizations

The purities and crystal structures of the samples were investigated by X-ray diffraction (XRD,

GBC, MMA) using Cu Kα radiation. The sample morphology and elemental analysis were

performed by field emission scanning electron microscopy (FESEM, JEOL-7500). The

composition of the samples was examined by an energy dispersive spectroscopy (EDS) instrument

coupled to the FESEM. The valences of elements were studied by X-ray photoelectron

Page 86: P-block-based Ferroelectric-photocatalyst compounds

73

spectroscopy (XPS, PHI660), which was performed using a monochromatic Cu Kα X-ray source.

Morphologies and structures were further examined by transmission electron microscopy (TEM)

and selected-area electron diffraction (SAED) (JEOL, JEM-2010). An ultraviolet-visible

spectrophotometer (UV-Vis, Shimadzu-3600) was used to collected diffuse reflectance spectra

(DRS) with an integrating sphere attachment to the instrument, with BaSO4 providing a

background between 200 nm and 800 nm. The on-off photocurrent–time-response (bias voltage =

0 V) and the Mott–Schottky curves were collected with a VSP-300 potentiostat, and a 300 W Xe

lamp was used as a simulated light source. The samples (BSO-HNS, BSO1, BSO2), Pt wire,

Ag/AgCl, and Na2SO4 solution (1 M) were used as working, counter, and reference electrodes, and

as the electrolyte, respectively. The working electrode was prepared via the dip-coating method on

indium tin oxide (ITO) and had an area 1 cm2 in size. The Mott–Schottky measurements were

performed at the frequency of 5 kHz. The photoluminescence (PL) spectra was carried out on a

Hitachi F-4500 fluorescence spectrophotometer at room temperature and obtained with excitation

wavelength at 320 nm. The ferroelectric properties of BSO-HNS were characterized by a

piezoresponse force microscopy (PFM) with a conductive Pt/Ir-coated silicon tip (Type: EFM,

Nanoworld. Force constant: 2.8 N/m). For the switching spectroscopy PFM (SS-PFM)

measurements, a sweeping DC bias (Frequency = 0.2 Hz) in the range of ± 25 V was applied to

provide BSO-HNS nanosheets with the essential electric field, and the quality factor (Q) was 288

(i.e., the gain in a weak PFM signal). During acquisition of a single curve, five cycles of sweeping

triangle/square waves were applied to show reproducibility. Igor Pro 6.36 Software was used to

obtain the butterfly loops and analyse the data. In each step, the DC bias was switched ON and

OFF to minimize electrostatic forces, and the data was recorded while the DC bias was OFF,

ensuring that the deformation of the sample was entirely due to the ferroelectric effect.

5.2.3 Photocatalytic activity evaluation

The photocatalytic activity was evaluated by the degradation of Rhodamine B (RhB) and phenol

(Sigma-Aldrich), which were used as model organics pollutants in the photocatalytic degradation

experiment at room temperature. A 300 W Xe lamp was used as the simulated light source for

photocatalytic performance. In a typical process, 50 mL of RhB (10 mg/l) (for phenol 20 mg/l),

and 0.05 g of the as-prepared sample were added together in a 100 mL beaker with a diameter of 6

cm. The suspension was magnetically stirred in the dark for 30 min to reach the adsorption-

desorption equilibrium, and then the mixture was irradiated with the light source. At certain time

intervals, a 3 mL aliquot was taken and centrifuged to remove the photocatalyst. The

photodegradation activity was analysed using the UV-Vis spectrophotometer by recording the

variation of the peak absorption 556 nm and 269 nm for RhB and phenol respectively. To

calculate the reaction kinetics Kt of RhB and phenol degradation, the pseudo-first-order kinetic

model was derived.

5.2.4 Theoretical calculation

All density functional theory (DFT) calculations were performed using the Vienna Ab Initio

Simulation Package (VASP) package. Generalized gradient approximation (GGA) was applied to

Page 87: P-block-based Ferroelectric-photocatalyst compounds

74

treat the exchange-correlation energy with the Perdew−Burke−Ernzerhof (PBE) functional and the

projector augmented wave (PAW) method was employed to describe electron−ion interactions.

The cut-off energy was set as 400 eV. Structures were fully relaxed until the convergence

tolerance of force on each atom was smaller than 0.01eV. The energy converge criteria was set to

be 1×10-5 eV for self-consistent calculations and k-point sampling was 5×7×7. Isosurface is 0.01

eV/Å3.

5.3 Results and discussion

The structure and element composition characterizations of the as-prepared samples were

carried out by XRD and XPS. As shown in Figures 5-1,5-2 and 5-3, all the recorded peaks of

BSO-HNS demonstrate the coexistence of two phases, the orthorhombic Bi2SiO5 phase (JCPDS:

36-0287) and the cubic Bi4(SiO4)3 phase (JCPDS: 002-0488). The pure phases BSO1 and BSO2

are also indexed to Bi2SiO5 phase (JCPDS: 36-0287) and Bi4(SiO4)3 phase (JCPDS: 002-0488),

respectively. No other peaks from possible impurities were detected. To confirm the absence of

any surface contamination materials, in particularly Br, the surface chemical composition of the

sample was analyzed by XPS, as illustrated in Figure 5-4 a, where only the elements Bi, Si, O, and

C were observed in the survey scan spectrum. For this heterostructure, the main peaks located at

158.9 and 164.3 eV correspond to Bi 4f7/2 and Bi 4f5/2, respectively, whereas the peaks of Si 2p

and O 1s are located at 101 and 530 eV respectively. EDS analysis of the BSO-HNS sample

(Figure 5-4 b) also confirmed the presence of Bi, Si, O, and no other impurity elements have been

detected. (The element carbon was from the background.)

The morphology and heterostructure features of the sample were investigated and confirmed by

FESEM and TEM. Flake-like nanostructures with a thickness of 10-20 nm can be observed from

the SEM image of the as-prepared BSO-HNS sample, as shown in Figure 5-5.

Figure 5-1 XRD pattern of as-prepared BSO-HNS

Page 88: P-block-based Ferroelectric-photocatalyst compounds

75

Figure 5-2 XRD pattern of as-prepared BSO1

Figure 5-3 XRD pattern of as-prepared BSO2

Page 89: P-block-based Ferroelectric-photocatalyst compounds

76

Figure 5-4 (a) XPS survey spectrum and (b) EDS of as-prepared BSO-HNS.

The as-prepared BSO-HNS sample with two phases shows flake structures with a thickness of 10 -

20 nm. The TEM and high-resolution TEM (HRTEM) images (Figure 5-6) show that some tiny

nanocrystals with sizes of 5-8 nm are inlaid in the flake matrix. The lattice fringes show d-spacing

of 0.274 nm for the flake matrix and 0.326 nm for the tiny nanocrystals, which can be assigned to

the d-spacing of the (020) planes for Bi2SiO5 and the d-spacing of (310) planes for Bi4(SiO4)3

respectively.

Figure 5-5 FESEM images of as-prepared BSO-HNS

Page 90: P-block-based Ferroelectric-photocatalyst compounds

77

Figure 5-6 TEM images of as-prepared BSO-HNS

The SAED pattern of the as-prepared BSO-HNS [inset image (1) in Figure 5-7] further confirms

the presence of Bi2SiO5 and Bi4(SiO4)3. As a layered material, the side view HRTEM of the

Bi2SiO5 sheet presents a clear layer structure with a layer spacing of 0.76 nm which is

corresponding to (200) planes (see the inset image (2) in Figure 5-7). In addition, in theoretically,

the angle between (200) and (020) planes is 90°, therefore, these results indicate that the top

surface sheet is dominated by (100) facet.

Figure 5-7 HRTEM images, inset (1): the SAED pattern, inset (2):

HRTEM image side view of as-prepared BSO-HNS.

Page 91: P-block-based Ferroelectric-photocatalyst compounds

78

Figure 5-8 the top schematic is atomic configuration in the top

view of (100) facets and the bottom schematic is the crystal

orientation and polarization direction of as-prepared BSO-HNS.

Depending on above explanation, the crystal orientation, polarization direction and atomic

configuration in the top view of (100) facets of as-prepared BSO-HNS have been determined as

shown in Figure 5-8. The Bi4(SiO4)3 phase is observed to resemble as dots that act as islands on

the Bi2SiO5 surface, which demonstrates clear contacted at the interfaces as well as good

compatibility of the lattices, indicating the formation of heterogeneous nanostructures. It is

proposed that, in the one-pot hydrothermal reaction, this large 2D flat exposed surface of Bi2SiO5

matrix grown via the crystallization of Bi2O22+ and SiO3

2- provides a good platform for the growth

of the 0D Bi4(SiO4)3 dots by the nucleation of Bi3+ and SiO44-.

Bi2SiO5 has been reported to be a 2D ferroelectric compound at room temperature in previous

studies [22, 23]. For the sheet-like nanostructures, the main direction of the polarization and

internal electric field is parallel (in-plane) to the 2D exposed surface, whereas the weak direction

of the polarization and internal electric field is perpendicular (out-of-plane) to the 2D exposed

surface, which is expected to facilitate the quick diffusion of the photoinduced carries to the large

exposed surface. To evaluate the ferroelectric properties of BSO-HNS, SS-PFM measurements

were carried out to study the polarization in the micro-size nanosheets. Figure 5-9 a and b shows

atomic force microscopy (AFM) topography and PFM amplitude images of the BSO-HNS sample,

showing flake-like morphology with the flakes having a lateral dimensional size of a few microns

and a thickness of 15 nm and the PFM image shows a the upward-polarized mono-domain

structure and zero PFM amplitude at the boundary of the flakes. The SS-PFM measurements were

carried on the spots indicated in Figure 5-9 band 5-10 a and b show the macroscopic hysteresis

loops, which resemble butterfly loops, and the corresponding phase switching curves. The

hysteresis amplitude − DC voltage loops are evidence of ferroelectric behavior with 180°

Page 92: P-block-based Ferroelectric-photocatalyst compounds

79

switching of the phase. Minimal deformation was exhibited at the phase-switching DC bias, which

means that responses from the nascent domains generated by the tip bias and the surrounding

material are compensating each other [27].

Figure 5-9 SS-PFM results on the BSO-HNS sample: (a) AFM topography with a white line

cutting cross-section the flake profile with a step height of 15 nm; image scale: 5 × 5 μm2, (b)

PFM amplitude image (white letters a and b represent the flake and substrate area respectively.

The deformation response and phase switching were both stable within the five cycles of

measurement, indicating the reversible nature of the ferroelectric polarization. Non-symmetric

coercive voltages were observed in range from 10 to -20 V in amplitude (Figure 5-10 a and b),

corresponding to the non-symmetric coercive voltages in the phase, since the electrostatic

contribution between the tip and the surface of the sample can lead to asymmetry in the hysteresis

loop curves [28]. The piezoelectric coefficient (dzz) for the out-of-plane direction was calculated

from the relationship between the amplitude (L) (i.e. the surface displacement or deformation) and

the AC applied voltage, Vac. dzz can be obtained from following equation [29]: dzz = L/Q∙Vac where

Q is the quality factor that describes the gain in a weak PFM signal in the system (Asylum

Research), Vac= 200 mV and L = 500 pm on average at 10 DC bias which calculated from

hysteresis loop. The obtained piezoelectric coefficient dij is 8.7 pm/V, which is a slightly lower

than the dzz of BSO bulk (11 pm/V out-plane) in Ref. [30]. The low dzz value is attributed to less

thickness, which, in turn, leads to low effective domains. For comparison, the same measurement

was also conducted on the substrate. The amplitude loops only showed noise, and no phase

switching was observed, which confirmed the butterfly loop generated from the ferroelectric

sample (Figure 5-11 a and b). Based on the above discussion, it is confirmed that the BSO-HNS

heterostructures remain ferroelectric in the nanoscale.

Page 93: P-block-based Ferroelectric-photocatalyst compounds

80

Figure 5-10 (a) PFM amplitude and (b) phase hysteresis loops for 5 cycles of the BSO-HNS sample.

Figure 5-11 PFM amplitude and (B) phase curves of as-prepared BSO-HNS for 5 cycles of

sweeping triangle/square waves on the substrate, the measurement was also conducted on the

substrate.

The diffuse reflectance spectra of the BSO-HNS, BSO1, and BSO2 samples were collected to

evaluate the optical properties of the as-prepared samples. According to their spectra (Figure 5-

12), all three samples have light absorption capability in the UV light region. The band gaps of the

samples were estimated from the onsets of the absorption edges by using the Tauc formula [31].

Page 94: P-block-based Ferroelectric-photocatalyst compounds

81

Figure 5-12 UV-Vis diffuse reflectance spectra of BSO-HNS,

BSO1, and BSO2 with the inset at the top right showing the

derivation of the band-gap values of the as-prepared samples

derived from the diffuse reflectance spectra.

As shown in the upper right inset image of Figure 5-11, the calculated band gaps are 3.5 eV, 3.3

eV, and 3.0 eV for BSO1, BSO-HNS, and BSO2, respectively. Obviously, the as-prepared BSO-

HNS sample with 0D Bi4(SiO4)3 dots on the Bi2SiO5 surface possesses stronger light absorption

capability than pure Bi2SiO5.

The photocatalytic activity of the BSO-HNS sample, as well as the BSO1and BSO2 samples,

was evaluated by degrading organic compounds under simulated solar light. RhB dye and phenol

were used here as a colorful dye pollutant and a typical colorless contaminant, respectively.

Before photocatalytic reaction, adsorption performance of three samples BSO1, BSO-HNS and

BSO2 have carried out in the dark in the interval 30 min for each RhB and phenol. The three

samples were compared as shown in Figure 5-13. From Figure 5-13 the adsorption performance of

the three samples reached their saturation within 30 min. It is clearly shown that BSO1 has

maximum adsorption for RhB and then BSO2 and BSO-HNS respectively, whereas the lower

adsorption performance has shown in phenol in three samples. For the photocatalytic degradation

test, as shown in Figure 5-14 and 5-15. The BSO-HNS sample exhibited the best efficiency toward

the photodegradation of RhB. The concentration of RhB relative to the absorption intensity at 556

nm decreased rapidly with increasing irradiation time and nearly reached zero within 50 min. In

contrast, only 30% and 15% of the RhB were decomposed under irradiation for 50 mins over the

BSO1 and BSO2 samples, respectively.

Page 95: P-block-based Ferroelectric-photocatalyst compounds

82

Figure 5-13 Adsorption performance of three samples BSO1, BSO-

HNS and BSO2 has carried out in dark in interval 30 min for each

RhB and phenol.

Figure 5-14 UV-Vis absorbance spectra of the photodegradation of

RhB dye by BSO-HNS under UV-Vis light, recorded after different

degradation times

To further explain the reaction kinetics of photocatalytic degradation, the pseudo-first-order

linear relationship is shown in the inset image of Figure 5-15. The constant rate k is 0.064 min-1

for BSO-HNS, which is much higher than those of BSO1 (0.0089 min-1) and BSO2 (0.0032 min-

1). This suggests that the photocatalytic activity of the 0D-2D heterostructures is about seven times

higher than for either of the pure phase samples. The photodegradation of phenol reached 35% in

the presence of the BSO-HNS sample after 150 mins under simulated solar light irradiation, as

shown in Figure 5-16, which demonstrates this sample’s good photocatalytic activity towards the

degradation of a colorless contaminant.

Page 96: P-block-based Ferroelectric-photocatalyst compounds

83

Figure 5-15 photodegradation rate of RhB dye by BSO-HNS, BSO1,

and BSO2 under UV-Vis light, with the inset showing a kinetic study

of the photocatalytic degradation

The photocatalytic durability of the BSO-HNS sample was evaluated by recycling the sample

after photodegradation. After each cycle, the catalyst was collected by centrifugation, as shown in

Figure 5-17 a. The decline in the photocatalytic activity was only 3.7% after 5 cycles over more

than 4 hours, indicating the high stability of the photocatalytic activity of BSO-HNS. To further

evaluate the chemical stability of the catalyst, XRD and XPS analysis of the BSO-HNS sample

after five cycles of photocatalytic reaction were carried out. As shown in Figure 5-17 b, all the

main peaks for Bi2SiO5 and Bi4(SiO4)3 remain. The XPS results (Figure 5-18) show no obvious

shift for all the characteristic peaks, indicating no valence change for the elements in this

composite, which indicates the good stability of this heterogeneous system.

Page 97: P-block-based Ferroelectric-photocatalyst compounds

84

Figure 5-16 photodegradation rate of phenol by BSO-HNS under

UV-Vis light, with the inset showing the UV-Vis absorbance spectra

of the degradation of phenol.

Figure 5-17 5 recycling runs to test the degradation of RhB by BSO-HNS under UV-

Vis light; (b) XRD patterns before and after 5 cycling runs to test the

photodegradation of RhB by BSO-HNS under UV-Vis light.

In comparison with the pure Bi2SiO5 and Bi4(SiO4)3 samples, the BSO-HNS sample shows a

significant enhancement in its photocatalytic activity, which may attribute to the more efficient

separation of the photoexcited electron-hole pairs in the as-prepared 0D-2D heterogeneous

structure.

Page 98: P-block-based Ferroelectric-photocatalyst compounds

85

Figure 5-18 XPS spectra of Bi 4f before and after 5 recycling

runs in the photodegradation of RhB by BSO-HNS under UV-

Vis light.

To verify this, as shown in Figure 5-19, the photocurrent response of the different samples was

investigated under intermittent light illumination (same wavelength range as that used in the

photocatalytic reactions). It can be clearly seen that the photocurrent for the BSO-HNS electrode

increased/decreased rapidly as the light was turned on/off. The transient photocurrent responses of

BSO-HNS electrode were quite stable and higher by three times than those of BSO1 and BSO2

electrodes. It is well known that enhancement in photocurrent response originates from improved

photoinduced charge separation. In view of the negative photocurrent responses produced under

the bias applied to the working electrodes in the electrochemical system here, the photocurrent

was produced mainly by the diffusion of the photogenerated holes to the ITO. Meanwhile, the

electron acceptor in the electrolyte took up the photoinduced electrons. Therefore, the enhanced

transient photocurrent response of BSO-HNS indicates a more efficient separation of the

photoinduced electron-hole pairs or a longer lifetime of the photogenerated charge carriers as

compared to the BSO1 and BSO2.

To further confirmation of charge separation, the PL spectra were carried out with excitation

light 320 nm. The PL results of BSO-HNS and BSO2 samples as shown in Figure 5-20, the higher

intensity emission spectra of BSO2 compare with BSO-HNS, indicating that BSO-HNS has higher

charge separation compare with BSO2.

Page 99: P-block-based Ferroelectric-photocatalyst compounds

86

Figure 5-19 Photocurrent density response under UV-

vis light.

The synergistic effects of high polarization in the 2D ferroelectric Bi2SiO5 nanosheets and the

band bending at the 0D-2D interfaces of the heterostructures is believed to be the most important

factor behind the enhanced photocurrent response and photocatalytic performance of the BSO-

HNS sample. To reveal the mechanism of the enhanced photocurrent response and improved

photocatalytic activity of BSO-HNS, the band-structure properties and ferroelectric properties of

the samples were investigated.

Figure 5-20 PL spectra for each BSO2 and BSO-HNS

samples.

The band positions including the conduction band (ECB) and the valence band (EVB) for BSO-

HNS, BSO1, and BSO2 were examined by Mott-Schottky measurements [32], as shown in Figure

5-21. The flat-band potentials (Efb) of BSO-HNS, BSO1, BSO2 are estimated to be −0.96, -0.92

and -0.94 V (vs. Ag/AgCl), respectively, or -0.76, -0.72, and -0.74 V (vs. normal hydrogen

electrode (NHE)) by Mott–Schottky analysis. These negative values and the positive slope

indicate n-type semiconducting properties for these three samples, Efb is considered to be located

Page 100: P-block-based Ferroelectric-photocatalyst compounds

87

just under the conduction band. Thus, based on the calculated band gaps mentioned above, a band

diagram of the BSO-HNS heterostructure with the valence band and conduction band edge

positions was constructed. As shown in Figure 5-22, the band bending between Bi2SiO5 and

Bi4(SiO3)3 in this system transfers the photoinduced electrons in the 0D Bi4(SiO3)3 to the surface

of the 2D Bi2SiO5, and the holes are transferred from 2D Bi2SiO5 to Bi4(SiO3)3.

Figure 5-21 Mott–Schottky plots of BSO-HNS, BSO1,

and BSO2.

This heterostructure results in high charge-separation efficiency in the localized areas of the 0D-

2D interface, which would not only significantly improve the separation/diffusion rate of the

photogenerated charge carriers, but also strongly affects the polarization of the 2D Bi2SiO5 matrix.

For further photocatalytic mechanism, it has been proven in literature that internal electric field

(IEF) (include spontaneous polarization) can promote charge separation [1, 3, 5, 33], particularly

Bi-based photocatalyst [34], and the IEF direction in nanosheet has shown different photocatalytic

activity. According to literature, the IEF perpendicular to the [Bi2O2]2+ slabs (i.e. perpendicular on

c-axis) has more photocatalytic activity than IEF generated parallel along [Bi2O2]2+ slabs because

shorter diffusion distance of charges to reach the surface [35-37].

Page 101: P-block-based Ferroelectric-photocatalyst compounds

88

Figure 5-22 Schematic diagram of the redox potentials of, BSO1, BSO2 and Possible photocatalytic

mechanism of the BSO-HNS sample according to the

Mott–Schottky and DRS spectroscopy results.

In our sample, it has been reported that IEF is perpendicular to the [Bi2O2]2+ slabs as well as

main IEF generated parallel along [Bi2O2]2+. Therefore, we carried out theoretical calculation as

shown in Figure 5-23 to confirm polarization on (100) facet which it is believed that polarization

can promote charge separation. This result suggests that the spontaneous electric polarization in

Bi2SiO5 is attributable to misaligned positive and negative charge centers, it was confirmed [22]

there are out-plan and in-plan polarization in Bi2SiO5 and each polarization direction has a role in

charge separation. Beside IEF, the heterostructure based IEF have shown higher charge separation

[38, 39]. The photo-generated charge electron-hole pairs which separated by out-plan polarization

can reach surface and interface surface where there is the band bending at the 0D-2D interface of

the heterostructure [38].

Figure 5-23 Charge density map of (100) facet of

BSO-HNS

Some holes would be injected to lower positive potential of VB of Bi4Si3O12 in the interface

Page 102: P-block-based Ferroelectric-photocatalyst compounds

89

region. Consequently, the charge electron-hole pairs have been separated. On the other hand, the

excited electron in Bi4Si3O12 could be injected to Bi2SiO5 surface and move away from Bi4Si3O12

due to effective in-plan polarization in (100) facet [39]. However, more explorations are needed in

future to further clarify photocatalytic mechanism because of the complexity of interfacing

semiconductor with ferroelectric materials.

5.4 Conclusion

In summary, 0D–2D bismuth silicate based heterostructures were successfully synthesized by the

in situ growth of 0D Bi4(SiO4)3 dots on 2D Bi2SiO5 nanosheets via a one-pot hydrothermal

process. The heterogeneous photocatalyst exhibits much enhanced photocatalytic activity towards

the degradation of RhB under simulated solar light irradiation, which is seven times higher than

for either of the pure phase samples. The enhanced activity is attributed to the effective separation

of electron–hole pairs due to the synergistic effects of polarization in the ferroelectric Bi2SiO5

nanosheets and the band bending at the 0D–2D interface of the heterostructures. Such a composite

photocatalyst is promising for water purification applications and environmental remediation.

5.5 References

1. Khan, M.A., M.A. Nadeem, and H. Idriss, Ferroelectric polarization effect on surface chemistry

and photo-catalytic activity: A review. Surface Science Reports, 2016. 71(1): p. 1-31.

2. Li, L., P.A. Salvador, and G.S. Rohrer, Photocatalysts with internal electric fields. Nanoscale,

2014. 6(1): p. 24-42.

3. Lou, Z., P. Wang, B. Huang, Y. Dai, X. Qin, X. Zhang, Z. Wang, and Y. Liu, Enhancing Charge

Separation in Photocatalysts with Internal Polar Electric Fields. ChemPhotoChem, 2017. 1(5):

p. 136-147.

4. Zhang, Q., Z. Dai, G. Cheng, Y. Liu, and R. Chen, In-situ room-temperature synthesis of

amorphous/crystalline contact Bi2S3/Bi2WO6 heterostructures for improved photocatalytic

ability. Ceramics International, 2017. 43(14): p. 11296-11304. 5. Li, H., Y. Sang, S. Chang, X. Huang, Y. Zhang, R. Yang, H. Jiang, H. Liu, and Z.L. Wang,

Enhanced Ferroelectric-Nanocrystal-Based Hybrid Photocatalysis by Ultrasonic-Wave-

Generated Piezophototronic Effect. Nano Letters, 2015. 15(4): p. 2372-2379.

6. Cui, Y., J. Briscoe, and S. Dunn, Effect of Ferroelectricity on Solar-Light-Driven Photocatalytic

Activity of BaTiO3—Influence on the Carrier Separation and Stern Layer Formation. Chemistry

of Materials, 2013. 25(21): p. 4215-4223.

7. Park, S., C.W. Lee, M.-G. Kang, S. Kim, H.J. Kim, J.E. Kwon, S.Y. Park, C.-Y. Kang, K.S.

Hong, and K.T. Nam, A ferroelectric photocatalyst for enhancing hydrogen evolution: polarized

particulate suspension. Physical Chemistry Chemical Physics, 2014. 16(22): p. 10408-10413.

8. Lou, Z., B. Huang, Z. Wang, X. Ma, R. Zhang, X. Zhang, X. Qin, Y. Dai, and M.-H. Whangbo,

Ag6Si2O7: a Silicate Photocatalyst for the Visible Region. Chem. Mater., 2014. 26(13): p. 3873-

3875. 9. Zhu, X., Z. Wang, B. Huang, W. Wei, Y. Dai, X. Zhang, and X. Qin, Synthesis of Ag9(SiO4)2NO3

through a reactive flux method and its visible-light photocatalytic performances. APL Mater.,

2015. 3(10): p. 104413.

10. Lou, Z., Z. Wang, B. Huang, Y. Dai, G. Wang, Z. Jiang, X. Zhang, X. Qin, and Y. Li, One-Step

Synthesis of Amorphous Silver Silicates with Tunable Light Absorption Spectra and

Photocatalytic Activities in the Visible Region. Chemistry – A European Journal, 2015. 21(24):

p. 8706-8710.

11. Al-keisy, A., L. Ren, D. Cui, Z. Xu, X. Xu, X. Su, W. Hao, S.X. Dou, and Y. Du, A ferroelectric

photocatalyst Ag10Si4O13 with visible-light photooxidation properties. Journal of Materials

Chemistry A, 2016. 4(28): p. 10992-10999.

Page 103: P-block-based Ferroelectric-photocatalyst compounds

90

12. Shang, J., W. Hao, X. Lv, T. Wang, X. Wang, Y. Du, S. Dou, T. Xie, D. Wang, and J. Wang,

Bismuth Oxybromide with Reasonable Photocatalytic Reduction Activity under Visible Light.

ACS Catal., 2014. 4(3): p. 954-961.

13. Feng, H., Y. Du, C. Wang, and W. Hao, Efficient visible-light photocatalysts by constructing

dispersive energy band with anisotropic p and s-p hybridization states. Current Opinion in

Green and Sustainable Chemistry, 2017. 6: p. 93-100.

14. Li, H., J. Shang, Z. Ai, and L. Zhang, Efficient Visible Light Nitrogen Fixation with BiOBr

Nanosheets of Oxygen Vacancies on the Exposed {001} Facets. Journal of the American

Chemical Society, 2015. 137(19): p. 6393-6399.

15. Tian, F., H. Zhao, G. Li, Z. Dai, Y. Liu, and R. Chen, Modification with Metallic Bismuth as Efficient Strategy for the Promotion of Photocatalysis: The Case of Bismuth Phosphate.

ChemSusChem, 2016. 9(13): p. 1579-1585.

16. Zhang, L., W. Wang, S. Sun, J. Xu, M. Shang, and J. Ren, Hybrid Bi2SiO5 mesoporous

microspheres with light response for environment decontamination. Applied Catalysis B:

Environmental, 2010. 100(1–2): p. 97-101.

17. Chen, R., J. Bi, L. Wu, W. Wang, Z. Li, and X. Fu, Template-Free Hydrothermal Synthesis and

Photocatalytic Performances of Novel Bi2SiO5 Nanosheets. Inorganic Chemistry, 2009. 48(19):

p. 9072-9076.

18. Zanardi, S., A. Carati, G. Cruciani, G. Bellussi, R. Millini, and C. Rizzo, Synthesis,

characterization and crystal structure of new microporous bismuth silicates. Microporous and

Mesoporous Materials, 2006. 97(1–3): p. 34-41. 19. Liu, X., W. Wang, Y. Liu, B. Huang, Y. Dai, X. Qin, and X. Zhang, In situ synthesis of

Bi2S3/Bi2SiO5 heterojunction photocatalysts with enhanced visible light photocatalytic activity.

RSC Advances, 2015. 5(69): p. 55957-55963.

20. Di, J., J. Xia, Y. Huang, M. Ji, W. Fan, Z. Chen, and H. Li, Constructing carbon quantum

dots/Bi2SiO5 ultrathin nanosheets with enhanced photocatalytic activity and mechanism

investigation. Chemical Engineering Journal, 2016. 302: p. 334-343.

21. Cheng, G., J. Xiong, H. Yang, Z. Lu, and R. Chen, Facile solvothermal synthesis of uniform

sponge-like Bi2SiO5 hierarchical nanostructure and its application in Cr(VI) removal. Materials

Letters, 2012. 77: p. 25-28.

22. Taniguchi, H., A. Kuwabara, J. Kim, Y. Kim, H. Moriwake, S. Kim, T. Hoshiyama, T. Koyama,

S. Mori, M. Takata, H. Hosono, Y. Inaguma, and M. Itoh, Ferroelectricity Driven by Twisting of Silicate Tetrahedral Chains. Angew. Chem., Int. Ed., 2013. 52(31): p. 8088-8092.

23. Park, J., B.G. Kim, S. Mori, and T. Oguchi, Tetrahedral tilting and ferroelectricity in Bi2AO5

(A=Si, Ge) from first principles calculations. Journal of Solid State Chemistry, 2016. 235: p. 68-

75.

24. Low, J., J. Yu, M. Jaroniec, S. Wageh, and A.A. Al-Ghamdi, Heterojunction Photocatalysts.

Advanced Materials, 2017: p. 1601694-n/a.

25. Dai, Z., F. Qin, H. Zhao, F. Tian, Y. Liu, and R. Chen, Time-dependent evolution of the

Bi3.64Mo0.36O6.55/Bi2MoO6 heterostructure for enhanced photocatalytic activity via the interfacial

hole migration. Nanoscale, 2015. 7(28): p. 11991-11999.

26. Ding, J., Z. Dai, F. Qin, H. Zhao, S. Zhao, and R. Chen, Z-scheme BiO1-xBr/Bi2O2CO3

photocatalyst with rich oxygen vacancy as electron mediator for highly efficient degradation of

antibiotics. Applied Catalysis B: Environmental, 2017. 205: p. 281-291. 27. Jesse, S., A.P. Baddorf, and S.V. Kalinin, Switching spectroscopy piezoresponse force

microscopy of ferroelectric materials. Applied Physics Letters, 2006. 88(6): p. 062908.

28. Alexe, N.G., A., Nanoscale Characterisation of Ferroelectric MaterialsScanning Probe

Microscopy Approach. 2005: p. 12-15.

29. Garcı́a, R. and R. Pérez, Dynamic atomic force microscopy methods. Surface Science Reports,

2002. 47(6–8): p. 197-301.

30. Seol, D., H. Taniguchi, J.-Y. Hwang, M. Itoh, H. Shin, S.W. Kim, and Y. Kim, Strong

anisotropy of ferroelectricity in lead-free bismuth silicate. Nanoscale, 2015. 7(27): p. 11561-

11565.

31. Tauc, J., R. Grigorovici, and A. Vancu, Optical Properties and Electronic Structure of

Amorphous Germanium. Phys. Status Solidi B, 1966. 15(2): p. 627-637. 32. Cardon, F. and W.P. Gomes, On the determination of the flat-band potential of a semiconductor

in contact with a metal or an electrolyte from the Mott-Schottky plot. J. Phys. D: Appl. Phys.,

1978. 11(4): p. L63.

Page 104: P-block-based Ferroelectric-photocatalyst compounds

91

33. Grinberg, I., D.V. West, M. Torres, G. Gou, D.M. Stein, L. Wu, G. Chen, E.M. Gallo, A.R.

Akbashev, P.K. Davies, J.E. Spanier, and A.M. Rappe, Perovskite oxides for visible-light-

absorbing ferroelectric and photovoltaic materials. Nature, 2013. 503(7477): p. 509-512.

34. Li, J., L. Cai, J. Shang, Y. Yu, and L. Zhang, Giant Enhancement of Internal Electric Field

Boosting Bulk Charge Separation for Photocatalysis. Advanced Materials, 2016. 28(21): p.

4059-4064.

35. Ye, L., X. Jin, X. Ji, C. Liu, Y. Su, H. Xie, and C. Liu, Facet-dependent photocatalytic reduction

of CO2 on BiOI nanosheets. Chemical Engineering Journal, 2016. 291(Supplement C): p. 39-46.

36. Li, J., L. Zhang, Y. Li, and Y. Yu, Synthesis and internal electric field dependent photoreactivity

of Bi3O4Cl single-crystalline nanosheets with high {001} facet exposure percentages. Nanoscale, 2014. 6(1): p. 167-171.

37. Jiang, J., K. Zhao, X. Xiao, and L. Zhang, Synthesis and Facet-Dependent Photoreactivity of

BiOCl Single-Crystalline Nanosheets. Journal of the American Chemical Society, 2012. 134(10):

p. 4473-4476.

38. Tan, C., G. Zhu, M. Hojamberdiev, K. Okada, J. Liang, X. Luo, P. Liu, and Y. Liu, Co3O4

nanoparticles-loaded BiOCl nanoplates with the dominant {001} facets: efficient

photodegradation of organic dyes under visible light. Applied Catalysis B: Environmental, 2014.

152(Supplement C): p. 425-436.

39. Li, Q., X. Zhao, J. Yang, C.-J. Jia, Z. Jin, and W. Fan, Exploring the effects of nanocrystal facet

orientations in g-C3N4/BiOCl heterostructures on photocatalytic performance. Nanoscale, 2015.

7(45): p. 18971-18983.

Page 105: P-block-based Ferroelectric-photocatalyst compounds

92

Chapter 6

Selective deposition of BiOI on the positive polar surface of

Bi4Ti3O12 by the polarization-adsorption effect and evaluation of

the photocatalytic performance

6.1 Introduction

The high recombination rates of photogenerated electron-hole pairs inside the bulk of a

photocatalyst, the back-reactions of intermediate species, and inactivity towards visible light are

regarded as the key issues blocking light-conversion efficiency. The driving force of charge

separation (DFCS) via an internal electric field (IEF) plays a key role in reducing recombination

rates [1-5]. The IEF that is in the p-n junction, unfortunately, only exists in the space charge layer

(due to band bending generated after contact between the n and p sides due to different Fermi

energy levels and different semiconductor types). Overall, according to a review on

heterojunctions, the enhancement of photocatalytic activity is not high [6]. The DFCS that is

provided by the ferroelectric photovoltaic effect (FPV) has demonstrated its ability to create high

photovoltage due to enhanced charge separation [7, 8]. This kind of nano-polycrystalline material

with disordered domains shows low photocatalyst activity, however, unless the IEF is controlled

(i.e. the polarization is orientated in the ferroelectric). It is confirmed that FPV produces spatially

separated oxidation/reduction (redox) sites on surfaces, and thus, it can reduce the back-reactions

on the surfaces and recombination [9]. To switch the polarization of a ferroelectric, an external

electric field is applied on the photoelectrode, therefore poling the polarization direction up and

down, according to the sign of the external electric field [10-13]. Nevertheless, control of

polarization has only been carried out in films. A ferroelectric-semiconductor (F-S) with a band

gap in the visible range, such as Bi4Ti3O12 (BTO) and BiFeO3 have been considered attractive

because they combine optoelectronic and ferroelectric properties [14-17]. Single crystal BTO

nanoplates prepared from molten salts have shown much higher photocatalytic activity with the

assistance of FPV [18-20]. The single crystal plate provides faceted surfaces, pathways for

electronic charge carriers and selective photodeposition, monodomain or longer disordered chains

in the crystal structure, and spatially separated redox positions on surfaces. Because of the BTO-

based layered structure, anisotropic behaviour is present, along with the presence of higher polarity

for the ab-plane than along the c-axis in BTO [16, 21]. The anisotropy in BTO is not only in

polarization, but also in the light absorption coefficient. It has been confirmed that the band gap in

BTO with (001) facet is a lower band gap than in the polycrystalline material [22]. The

polarization potential can be screened by charges which either come from the bulk of the

ferroelectric or from the environment (usually, a liquid, since the liquids contain many Ions,

cations, and polar molecules) and then generate a Stern layer causing charge separation. The

previous reports have studied the F-S/S interface to form junctions [11, 13]. Nevertheless, few

have studied F-S/S nanostructures in powder form with the growth controlled over the specific

Page 106: P-block-based Ferroelectric-photocatalyst compounds

93

polarization potential. In this work, we explore a new concept to form a heterojunction built-in

DFCS that has a positive polarization of BTO interfaced with an n-type semiconductor such as

BiOI in order to charge transfer. BiOI works as an electrons transporter for the electrons coming

from BTO due to its having a lower conduction band than BTO and a layered crystal structure

(including Bi2O2+2 slabs for each BiOI and BTO) and also a narrow band gap and easy growth.

The polarization-adsorption interaction is important for selective deposition on the surface. It has

been proven that the polar surface of a ferroelectric adsorbs ions and polar molecules from a

solution [23, 24], which then can undergo easy epitaxial growth as a photocatalyst [25]. Spatially

selective deposition under illumination can be also observed for the deposition of metals

(including Ag, Au, Pt, etc.) on the positively polarized boundaries of a ferroelectric surface [9],

and photocatalytic activity is significantly enhanced compared to deposition by a conventional

loading method where there is disorder in the interfacing materials. Moreover, photodeposition can

explore the pathways of electron-hole and redox position surfaces. We propose in this work that

this represents a new strategy to enhance photocatalytic activity and open up opportunities for

further exploration of ferroelectric-semiconductor/semiconductor interfaces.

6.2 Experimental

6.2.1 Synthesis of sample

The BTO plates were synthesised by the molten salt method. The BTO plates were prepared in as

in a previous reports by mixing bismuth oxide and titanium oxide with salts, but here, we used

little modification, and we prepared amorphous BTO powder by wet reaction of the raw materials

and then mixing them with salts. It was believed that this modification is easier than in previous

reports because it doesn’t need much mixing and requires a shorter time for reaction. In a typical

experimental process: Bismuth nitrate (Bi(NO3)3.5H2O, Sigma-Aldrich) and titanium butoxide

(C16H36O4Ti, Sigma-Aldrich) were used as raw materials. A typical synthesis for BTO plates is

described as follows: 4 mmol of Bi(NO3)3·5H2O was dissolved into a solution of 20 ml ethylene

glycol and 40 ml absolute ethanol, and then 3 mmol of C16H36O4Ti was dissolved in the above

solution drop by drop. The pH was adjusted by adding KOH up to pH = 10. The clear solution

changed to a white precursor under stirring, and it was kept under stirring for 4 h before being

filtered by centrifuge. The white powder was washed many times with water and ethanol.

Amorphous BTO was obtained after drying for a day. The weight ratio was 10 g : 1 g of the salts

(NaCl: KCl = 1:1) to the powder. The mixture was been mixed in a mortar for 5 min and then

sintered for 1 h at 750 °C and cooled naturally to room temperature. The product was washed

many times by water and ethanol to remove salts and then dried.

6.2.2 Selective photodeposition of metal/metal oxide on BTO plate:

Photoreduction was carried out at room temperature with no change in pH. A typical 1 ml of

AgNO3 (0.1 M AgNO3), which was used in the reduction process to deposit Ag, was added to a

BTO solution (0.1 g BTO was dispersed in 50 ml distilled water + 2 ml methanol). The suspension

was stirred for 0.5 h and then irradiated under a 300 W Xe lamp without an optical filter for 6 h.

After that the suspension was filtered and washed, and finally dried at 60 C. Photooxidation was

Page 107: P-block-based Ferroelectric-photocatalyst compounds

94

also carried out by the same procedure above, in which 1 ml of MnSO4 (0.1 M MnSO4) was used

for the oxidation process to deposit MnOx. It was added into a solution containing KIO3, and the

irradiation time was 0.5 h.

6.6.3 Selective deposition of BiOI on BTO plate:

Because of the polarity behaviour of the ferroelectric, 0.1 g of BTO powder was dispersed in 20 ml

ethylene glycol and then 150 ml of distilled water was added under sonication for 2 min, and then

the dispersion was stirred for 10 min to prepare the BTO suspension. Two solutions, one of

sodium iodide and a bismuth nitrate solution, were prepared separately. 1 mmol of NaI was

dissolved in 50 ml of distilled water, and Bi(NO3)3·5H2O was dissolved in 15 ml ethylene glycol,

followed by the addition of 35 ml water. Firstly, 0.25 ml of NaI solution was added to the BTO

suspension, which was then kept under stirring for 15 min before 0.25 ml of Bi(NO3)·5H2O

solution was added to the suspension and it was again kept under stirring for 15 min. The above

procedure was continued with the addition of NaI and Bi(NO3)3·5H2O until the amount became

0.75 ml for each. The colour of the suspension changed from white to light yellow, and this sample

was labelled as BTO-BI-1. For a greater BiOI concentration, more NaI and Bi(NO3)3·5H2O

solutions were added into the above suspension to achieve 1.25 ml and 1.75 ml of BiOI for the

next to samples, respectively, and the results were labelled as BTO-BI-2 and BTO-BI-3,

respectively. The colour of the samples (BTO-BI-1 and BTO-BI-2) was changed from light yellow

to yellow and yellow-orange, respectively. For comparison, the BTO-BI-3 was also prepared by a

conventional loading method. Typically, 1.75 ml NaI solution was added to the above suspension

and then 1.75 ml Bi(NO3)3·5H2O was added, which was previously solved in water, and it was

kept under stirring for 15 min and labelled BTO-BI-X. All samples were finally washed and then

dried. For pure BiOI, the same procedure above was followed without BTO powder.

6.6.4 Characterization

The purities and crystal structures of the samples were investigated by X-ray diffraction (XRD,

GBC, MMA) using Cu Kα radiation. The sample morphology was investigated and the elemental

analysis performed by field emission scanning electron microscopy (FESEM, JEOL-7500). The

composition of the samples was examined by using an energy dispersive spectroscopy (EDS)

instrument coupled to the FESEM. The valences of elements were studied by X-ray photoelectron

spectroscopy (XPS, PHI660), which was performed using a monochromatic Cu Kα X-ray source.

Morphologies and structures were further examined by transmission electron microscopy (TEM)

and selected-area electron diffraction (SAED) (JEOL, JEM-2010). An ultraviolet-visible

spectrophotometer (UV-Vis, Shimadzu-3600) was used to collect diffuse reflectance spectra

(DRS) with an integrating sphere attachment to the instrument, with BaSO4 providing a

background between 200 nm and 800 nm.

6.6.5 Thin film preparation

The thin films of BTO, BTO-BI-1, BTO-BI-2 and BTO-BI-3 were prepared via a metering rod on

fluorine-doped tin oxide (FTO) substrates that had an area of 2.5 cm2. The substrates were

previously cleaned with ethanol under sonication for 10 min. 0.005 g of powder was mixed with a

Page 108: P-block-based Ferroelectric-photocatalyst compounds

95

binder material such as 4-5 µl of N-methyl-2-pyrrolidone (NMP) and mixed in a mortar for 15 min

until the mixture became a paste. The metering rod method was used to prepare the thin films with

an area of 1 cm2, and they were then annealed at 150 °C for 3 h, which provided them with

mechanical stability. BTO thin films with positive poling, no poling, and negative poling were

prepared according to ref. [26] and then the polarization was switched in the opposite direction by

an applied external voltage, whereas the BiOI/BTO photoanodes were prepared by the growth of

BiOI on BTO thin film by the successive ionic layer adsorption and reaction (SILAR) method.

Electrochemical and photoelectrochemical measurements

The transient photocurrent, the photovoltage decay and the flat band potential were collected with

a VSP-300 potentiostat, and a 300 W Xe lamp was used with a UV-optical filter. The samples

(BTO, BTO-BI-1, BTO-BI-2, and BTO-BI-3), Pt wire, Ag/AgCl (KCl 3 M), and Na2SO4 solution

(0.1 M) were used as the working electrode, counter, and reference electrodes, and as the

electrolyte, respectively. The on-off light response was collected every 30 seconds by chopped

light for transient photocurrent measurements by chronoamperometry (CA) (bias voltage = 0 V),

whereas the photovoltage decay time was measured after light-off, using the open circuit voltage

technique (OCV). The flat band potential was measured according to The Mott– Schottky (MS)

method, which was performed at a frequency of 1 kHz in the dark at room temperature, and the

impedance was measured in the frequency range of 100 KHz to 0.05 Hz with applied voltage of 5

mV in the dark and under light, using electrochemical impedance spectroscopy (EIS).

6.6.6 Photocatalytic activity evaluation

The photocatalytic activity was evaluated by the degradation of Rhodamine B (RhB) and phenol

(Sigma-Aldrich), which were used as model organic pollutants in the photocatalytic degradation

experiment at room temperature. A 300 W Xe lamp with cut-off filter (λ > 420 nm) was used as

the light source for studying the photocatalytic performance. In a typical process, 25 mL of RhB

(10 mg l−1) (or phenol 20 mg l−1), and 0.025 g of as-prepared sample were added into a 100 mL

beaker with a diameter of 6 cm. The suspension was magnetically stirred in the dark for 30 min to

reach the adsorption-desorption equilibrium, and then the mixture was irradiated by the light

source. At certain time intervals, a 3 mL aliquot was taken and centrifuged to remove the

photocatalyst. The photodegradation activity was analysed using the UV-Vis spectrophotometer by

recording the variation of the peak absorption at 556 nm for RhB and 270 nm for phenol. For

comparison with other dyes, Methylene blue (MB) and Methylene orange (MO) were used for

comparison with RhB.

Page 109: P-block-based Ferroelectric-photocatalyst compounds

96

6.3 Results and Discussion

According to our procedure, the BTO powder was prepared using molten salts, and the colours

of the BTO-BI-1, BTO-BI-2, and BTO-BI-3 heterojunction samples were changed from milky to

milky-yellow, yellow, and yellow-orange, respectively on increasing the BiOI content.

Figure 6-1 (a) XRD patterns, (b) XPS survey spectra, (c, d, e, and f) high-resolution XPS spectra for Bi 4f, O 1s, I 3d, and Ti 2p, respectively, for the BTO, BiOI, BTO-BI-1, BTO-BI-2, and BTO-

BI-3 samples.

The sample characterizations were first conducted by X-ray diffraction (XRD) to analyse the

crystallinity and purity. The XRD patterns of the as-prepared samples, as shown in Figure 6-1(a),

indicate high crystallinity. All the diffraction peaks can be indexed to orthorhombic Bi4Ti3O12

(space group B2ab) and were without any observable impurities, consistent with the Joint

Committee on Powder Diffraction Standards (JCPDS) Card no. 01-089-7500. In the XRD pattern

of BTO-BI-3, the BiOI diffraction peaks are clearly observed which are 29.7° and 31.7°, assigned

to the (012) and (110) planes of tetragonal BiOI (JCPDS: 90-901-1785), whereas peaks of BiOI

were not observed in BTO-BI-1 and BTO-BI-2 because of the low content. Because the diffraction

peaks related to BiOI have not been demonstrated in XRD, except for BTO Bi-3, XPS was

conducted to confirm the chemical states and chemical compositions of the samples. The survey

Page 110: P-block-based Ferroelectric-photocatalyst compounds

97

spectrum for the BTO sample reveals that it only contains Bi, Ti, and O elements without any

impurity (Figure 6-1(b)) whereas in the heterojunctions, the samples include I element. The high-

resolution results demonstrate that the two peaks with binding energy at 164 and 158.6 eV

correspond to Bi 4f5/2 and Bi 4f7/2 of Bi+3 in the crystal (Figure 6-1(c)), respectively and there is a

slight shift in the binding energies in the heterojunctions. The binding energy was shifted to higher

binding energy toward pure BiOI with increasing BiOI content. The peak at 528.9 eV is assigned

to O 1s (Figure 6-1 (d)), and this peak could be deconvoluted into three peaks with binding

energies of 528.9, 529.2, and 531.7 , which correspond to Bi-O, Ti-O, and adsorbed hydroxyl

groups. The binding energies of the heterojunctions were shifted up. Moreover, a small peak at

155.5 eV corresponds to oxygen vacancy. Which is seen in each both pure BiOI and pure BTO.

The peaks at 630.3 eV and 618.8 eV are assigned to I 3d3/2 and I 3d5/2, respectively, and the

chemical state is I-1. There was no shift in the binding energy with decreasing BiOI content. The

binding energy peak at 457.1 eV is assigned to Ti 2p3/2, whereas the Ti 2p1/2 peak overlaps with Bi

4d3/2 at 462.8 eV with the chemical state Ti+4. The Ti 2p peaks were not clearly shifted with

increasing BiOI content. On comparing the XPS spectrum of pure BTO with that of a BTO/BiOI

heterojunction, the elements Bi and O were shifted up in binding energies with no contribution

from Ti and I, which could indicate that that Bi-O bonds of BTO interacted with the Bi-O bonds of

BiOI. They possess different electronegativity, and electrons transfer between them results in band

bending. The morphology of the samples was observed by FESEM and TEM images. The FESEM

image reveals that the as-prepared BTO sample consists of micron-sized plate-like particles with a

thickness of 50-60 nm (Figure 6-2(a)). The heterojunction samples BTO-BI-1, BTO-BI-2, and

BTO-BI-3 were also observed by FESEM. as shown in Figure 6-2(b), (c), and (d). It can be seen

that the BiOI has only grown at the edge of the BTO plates, and the growth of BiOI increased with

an increasing amount of NaI/Bi(NO3)3 ·5H2O solution in the reaction. In the BTO-BI-3 almost all

the edges of BTO appear to be covered, as shown in Figure 6-2(d), which was investigated by EDS

to find the elemental contents and distributions. Figure 6-2(f), (j), (d), (h), and (i) corresponds to

Bi, I, Ti, and O respectively. The TEM image shown in Figure 6-2(g) is more confirmation of the

morphology shown by the FESEM image. The single crystal nature of the Bi4Ti3O12 plates is

confirmed in the SAED pattern shown in Figure 6-2(k), and the top view surfaces of the BTO

plates are exposed (001) facets, which were determined from the theoretical interfacial angle of

90° between the (020) and (200) planes. The (004), (003), and (011) planes indicated by the yellow

lines in the Figure confirm the presence of BiOI in area enclosed by the yellow square in Figure 6-

2(g). For further confirmation, high resolution TEM (HRTEM) was conducted to observe the same

yellow area, and the BiOI shows (001) facets. The results for both BTO and BiOI indicate exposed

(001) facets. It should be noted that there is large spontaneous polarization along the a-axis due to

distortion of the entire TiO6 octahedral layer with respect to the Bi ions, and thus the edges of BTO

show polarity [3]. According to previous results on BTO and our results from HRTEM and XPS,

the mechanism for the growth BiOI on the edges of BTO could be explained. A physisorption

interaction exists between the terminal edge of BTO and the I-1 present with Bi+3 in solution.

Because of the permanent polarization of BTO, electrostatic interaction (van der Waals forces)

Page 111: P-block-based Ferroelectric-photocatalyst compounds

98

occurs between positively polarised BTO and the negative ions in solution, resulting in screening

of the positive polarization. Next, when the chemical force exists between the adsorbed substance

and the terminal edge of BTO in the presence of Bi+3, chemical reactions occur, forming chemical

bonds attaching BiOI to the BTO edge; the process is called chemisorption.

The photocatalytic activity of the samples was evaluated by decolouration of organic dye under

visible light. RhB dye was used as a colourful dye. Before the photocatalytic reaction, the

adsorption performance of all samples was investigated in the dark after an interval of 30 min. The

samples were compared, as shown in Figure 6-3(a), and it was found that the adsorption

performance of the samples reached saturation within 30 min and exhibited high adsorption

efficiency, possibly due to the polar surface of BTO. In addition, there are no significantly

different results between the samples in the adsorption process. The BiOI-BI-2 and BiOI-BI-3

samples exhibited the best efficiency toward the photodegradation of RhB, and because the BiOI-

BI-2 showed very close results to BiOI-BI-3, it has been neglected throughout the Figures. A

typical visible absorbance spectrum of RhB degradation by BTO-BI-3 is shown in Figure 6-3(b),

where the concentration of RhB is relative to the intensity of peak absorption at 556 nm, which

decreased rapidly with increasing irradiation time and nearly reached zero within 12 min (Figure

6-3(b)).

Figure 6-2 FESEM images of the (a) BTO,(b) BTO-BI-1, (c) BTO-BI-2, and (d) BTO-BI-3 plates,

respectively; (e) a typical single plate of BTO/BiOI; and the EDS mapping of (f) Bi, (g) I, (h) Ti,

and (i) O, and (g) TEM image for a BTO/BiOI heterojunction; (k) the corresponding SAED pattern,

and (l) the corresponding HRTEM image.

Page 112: P-block-based Ferroelectric-photocatalyst compounds

99

Page 113: P-block-based Ferroelectric-photocatalyst compounds

100

Figure 6-3 Photodegradation rates for the RhB dye by all samples under visible light. (b) Typical

visible absorbance spectra of the photodegradation of the RhB dye by BTO-BI-3 under visible

light, recorded after different degradation times. (c) Photodegradation rates of BTO, BTO-BI-1,

and BTO-BI-3 for phenol within 60 min. (d) Adsorption performance of BTO, BTO-BI-1, BTO-

BI-2, and BTO-BI-3 carried out in the dark after an interval of 30 min for RhB and MO. (e)

comparison of photodecolouration rate percentage of the RhB, MB, and MO dyes achieved by

BTO and BTO-BI-3 for different irradiation times under visible light. (f) Comparison of the

photodegradation by BTO-BI-3 and BTO-BI-X for RhB after 10 min.

Page 114: P-block-based Ferroelectric-photocatalyst compounds

101

The colourless and toxic pollutant phenol was used to evaluate the photodegradation

performance of the samples over 60 min, and the results demonstrated that 47% of phenol was

photodegraded during the 60 min by BTO-BI-3, as shown in Figure 6-3(c), which indicates that

BTO-BI-3 has excellent performance towards photodegradation of phenol. In contrast, only 12%

abd 20 % phenol degradation were achieved by BTO and BTO-BI-1, respectively, under visible

irradiation for 60 min. A comparison of the photodecolouration rate for other dyes with different

cations and anions, MB and MO dyes, respectively, was carried out using BTO and BTO-BI-3

with different irradiation times under visible light. The adsorption results showed (Figure 6-3(d))

excellent adsorption of RhB with adsorption capacity (Qe) of about 5 mg/g compared with 0.5

mg/g for MO. The photodegradation rates of cationic dyes (RhB, Crystal Violet (CV), and MB)

for the BTO and BTO-BI-3 samples were significantly higher than those for the anionic dyes (MO,

Orange G (OG). and Alizarin Yellow (AY)) (Figure 6-3(e). The surface polarity of the sample

could be behind the selective adsorption. The cationic dyes are highly adsorbed due to negative

potential of polarization in BTO because this group of dyes carry a positive charge in their

molecules when they are dissolved in water. Therefore, zeta-potential measurements (ζ) were

conducted in deionized water (pH = 7) at room temperature. The results revealed that BTO had a

negative value (ζ = -12 mV), indicating that BTO has negative surface. It is believed that

photodegradability of the negative polarization surface has higher activity than the positive

polarization due to significantly different photodegradation activity.

To further analyse the reason for the superior enhancement shown by BTO-BI-3. a comparison

has carried out between BTO-BI-3 that was grown by our method and BTO/BiOI prepared by the

conventional loading method with the same content of BiOI. The result, as shown in Figure 6-3(f),

is that BTO-BI-3 has much higher photodegradation performance than BTO-BI-X (conventional

Figure 6-4 (a-f) SEM image of BTO-BI-X sample, full spectrum EDS mapping of BTO-Bi-X

and EDS mapping of Ti, Bi, I, and O, respectively.

Page 115: P-block-based Ferroelectric-photocatalyst compounds

102

loading method) due to the randomly growth BiOI on BTO, as shown in Figure 6-4, which shows

the EDS mapping of the BTO-BI-X sample. The wrong type of deposition, for example, not whole

area deposition of the CB of BiOI on the CB of BTO or vice versa, cannot lead to enhancement in

charge separation. Beside this, the growth of BiOI on the BTO surface blocks the visible spectrum

for BTO.

The selective photodeposition of Ag on BTO has been carried out under UV-vis light to study

charge separation and polarization behaviour. The results as shown in Figure 6-5(a) confirm that

the positive potential edge due to polarization in BTO attracts electrons which are photogenerated

under light, and electrons transferred at the edge react with Ag+, resulting in the deposition of Ag

nanoparticles on the edge of BTO. Figure 6-5(b) and (c) are further confirmation of the identity of

Ag metal. In addition, the photooxidation process deposits MnOx on the top and bottom surfaces of

BTO. Thus, it can be confirmed that this ferroelectric material can separate photoinduced electrons

and holes and also possesses spatially separated oxidation/reduction sites.

Figure 6-5 (a) Scanning electron microscope (SEM) images of BTO plate with photodeposited

Ag on positive potential of polarization. (b) TEM image of deposited Ag on BTO. (c) SAED

pattern of the inside of the yellow rectangle in the TEM image in (b), and (d) SEM image of

photodeposited MnOx on BTO.

Page 116: P-block-based Ferroelectric-photocatalyst compounds

103

By macroscopically quantifying the amount of Ag:MnOx, it is possible to observe the reactivity of

photoreduction and the photooxidation rate. The amount of photodeposition of MnOx during over

0.5 h of reaction is higher than for Ag over 6 h of reaction, thus indicating that photooxidation is

much more reactive than photoreduction. It has confirmed by Kelvin probe force microscopy

(KPFM) in [18]with the same synthesis method and morphology that a higher positive potential

was found at the edge of BTO with difference voltage potential in dark and light is 23 mV.

The on-off photocurrent was used to investigate charge separation in the samples. The

photocurrent density was multiplied by at least six times because of the high charge separation of

BTO-BI-3 compared with BTO, as shown in Figure 6-6(a), and the photocurrent result is in

agreement with the photocatalytic activity. The electron recombination was further investigated by

photovoltage decay in the on-off light conditions. The decay time is related to the charge lifetime

and is considered as indicating the recombination rate level. The results are consistent with the

photocurrent results (Figure 6-6(b)), with BTO-BI-3 showing the slowest decay time, also

consistent with the photocurrent results.

To reveal the reason behind this enhancement, the band-structure properties of the samples,

including BiOI, need to be studied. The optical properties of the samples were investigated by the

diffuse reflectance. According to the absorption spectra (Figure 6-6(c)), all the samples show

visible light absorption, and then there is a shift to red light spectra from BTO to BTO-BI-3 on

increasing the BiOI content. The band gaps of the samples were estimated from the onsets of the

absorption edges by using the Tauc formula. As shown in the upper right inset image of Figure

6(c), the calculated band gaps are 3.05 eV, 2.54 eV, 2.1 eV and 2 eV for the BTO, BTO-BI-1,

BTO-BI-2, and BTO-BI-3 samples, respectively. The valence and conduction band positions were

determined by Mott-Schottky measurements. The flat-band potential (Efp) value is important for

determining the Fermi level, because the Efp has been considered nearly equal to the Fermi level.

For n-type semiconductors, the Fermi level is considered lower than the conduction band by 0.3 ̴

0.1 V. The Efp values of the BTO and BiOI films were estimated, as shown in Figure 6-6(d), to be -

0.66 and -0.59 V vs. Ag/AgCl (0.1 M KCl) corresponding to -0.372, -0.302 V vs. normal hydrogen

electrode (NHE), respectively. These positive slopes indicate n-type semiconducting properties of

these samples. Thus, the conduction bands are at -0.472 and -0.402 V and from the band gap

estimates (where the BiOI band gap was considered to be 1.8 eV according to previous reports),

the valence bands are 2.57 and 1.4 V, respectively. A typical BTO/BiOI heterojunction is useless

because the conduction band of BiOI is lower than that of BTO, and the valence band of BiOI is

lower than that of BTO, so that electrons and holes will collect in the BiOI, and thus, higher

charges recombination is expected. Herein, a single type of charge transfer from BTO to BiOI is

useful for charge separation, as shown in Figure 6-6 (d), the IEF that is provided by FPV is the

driving force that pushes the electrons to the edges of BTO plates. Because the BTO is n-type,

however, it is believed that some majority charge carriers (electrons) are driven to the positively

polarised regions of BTO, and because the edge of BTO, which is {110}, has a very narrow

surface with a small surface area (as shown in scanning electron microscope (SEM) and HRTEM

Page 117: P-block-based Ferroelectric-photocatalyst compounds

104

images), thus the positive pole of the polarization is roughly eliminated, resulting in low effective

positive polarization. It is thus expected exhibit low effective efficiency compared with the large

area of {001} of BiOI. Thus, BiOI provides a larger surface area with extra electron pathways.

Figure 6-6 (a) On-Off photocurrent and (b) photovoltage decay for samples under visible light.

(c) UV-vis diffuse reflectance spectra of BTO, BTO-BI-1, BTO-BI-2, and BTO-BI-3 plates, with

the inset at the top right of (c) the derivation of the band gap values of the four samples, which

are derived from the diffuse reflectance spectra. (d) MS plots of BTO and BiOI.

Page 118: P-block-based Ferroelectric-photocatalyst compounds

105

Furthermore, to further investigate the interfacing positive/negative poles of BTO with BiOI, a

BiOI/BTO photoanode with different switching polarization was prepared to measure

photocurrent, open circuit voltage (Voc), EIS spectra, and MS. The results are shown in Figure 6-6,

where there is higher photocurrent by positive poling than with the negative poling and no poling

photoanodes. Also, the positive poling led to a higher Voc than for no poling or negative poling.

EIS is important to investigate the charge transfer in the dark and under light, and the smaller

radius of arc in the spectrum indicate lower impedance and higher charge transfer. Thus, positive

poling in dark and under light demonstrated lower impedance. In the MS (Figure 6-7(d))

measurement, the flat band potential was demonstrated to be higher in the negative poling

photoanode than the no poling and positive poling photoanodes. The reason behind the different

properties could be that a barrier is naturally created in the interface for negative poling, so it is

believed that there is no charge transfer, even though there is no perfect polarization switching in

the ferroelectric materials based layered structure. Therefore, the photocurrent in negative poling is

nearly all in BiOI without any contribution of BTO photocurrent. Also, there is a high

photovoltage in the positive poling photoanode, as shown in Figure 6-7(a) and (b) because of the

long lifetime of photogenerated carriers due to charge separation. The higher impedance in the

Figure 6-7 (a) On-Off photocurrent and (b) open circuit photovoltage of samples. (c) Nyquist

plots at open-circuit potential under dark and illumination. (d) Mott–Schottky plots for positive

poling, no poling, and negative poling samples.

Page 119: P-block-based Ferroelectric-photocatalyst compounds

106

negative poling photoanode is due to the barrier in the interface region. It is believed that the

barrier was generated due to the creation of a virtual p-n junction. The impedance between BiOI

and the electrolyte is neglected because the same photoanode is used for positive and negative

poling.

The virtual p-type is due to accumulated photogenerated holes at the interface, and this thus gives

rise to the flat band potential, as shown in Figure 6-7(d). This kind of virtual p-n

junction with this band structure cannot contribute to charge separation, and the conduction band

(CB) of BiOI still has lower potential than the CB of BTO, since otherwise anodic photocurrent

can be observed. A schematic diagram has been drawn as shown in Figure 6-8 to illuminate the

band structures for positive poling and negative poling. To test the claim that holes play an

Figure 6-8 Schematic diagram of BTO/BiOI thin film and corresponding band structures for

negative poling and positive poling photoanodes.

Figure 6-9 (a) Bar chart to show role electrons and holes in photodegradation (b) SEM

image and EDS spectrum (inset) of BTO-BI-3 with photoreduction of Ag.

Page 120: P-block-based Ferroelectric-photocatalyst compounds

107

absolutely key role in the decolouration process for pure BTO, a scavenger effect was created by

adding sodium oxalate as a hole scavenger while added calcium iodate as an electron scavenger.

The bar plot, as shown in Figure 6-9(a) demonstrates that the photocatalytic activity toward RhB

photodegradation is

affected with added scavengers. Firstly, with the electron scavenger, the photocatalytic activity is

less affected with BTO compared with BTO-BI-3 demonstrating that holes are much more

effective toward the photodegradation of RhB in BTO, whereas the electrons play this role in

photodegradation with BTO-BI-3. Secondly, with the hole scavenger, the photocatalytic activity is

much more affected with BTO compared with BTO-BI-3, implying the role of electrons in BTO-

BI-3. As a result, the enhancement in photocatalytic activity of BTO-BI-3 occurs because of the

contribution of the electrons in the reduction process, while in BTO, only holes play a role in

photodegradation.

The high-performance activity of the heterojunction has a considerable effect on the band

structure at the interface, and because the BTO and BiOI have a layered crystal structure and close

band structures with a little higher CB of BTO than BiOI and only connection in the ab-plane

(BiOI) with exposed {001} facets, as shown in Figure 6-2(i), which mean continue bonding Bi-O

that it is possible for the pathways of electrons to cross from the Bi2O2 layer of BTO to the Bi2O2

layer of BiOI. Thus, the majority charge carriers (electrons) which are accumulated at the edge of

BTO due to the positive polarization are transferred from the edge of BTO to the CB of BiOI. The

transfer will stop when the Fermi levels are aligned. Consequently, salvation of the trapped

electrons in the positively polarised regions has been achieved, and thus pathways are created for

escape of the photoinduced electrons through BiOI. As can be seen in Figure 6-9 (b).

Although the charge transfer is single-type (electrons), but the photocatalytic activity is

significantly enhanced. The schematic diagram in Figure 6-10 shows BiOI in contact with the

Figure 6-10 Mechanism of BTO/BiOI heterojunction: (a) schematic illustration of growth of BiOI

on edge of BTO; (b) possible mechanism of heterojunction; (c) crystal structure for BTO and

BiOI shown at the interface; (d) theoretical calculation of interface band structure.

Page 121: P-block-based Ferroelectric-photocatalyst compounds

108

edges of BTO, and Figure 10 b shows the mechanism of the BTO/BiOI heterojunction, where it is

clearly observed that the photoinduced electrons at the positive edges of BTO could be easily

transferred to the CB of BiOI, and thus charge separation is achieved. Therefore, the photocatalyst

performance has been enhanced.

6.4 Conclusion

We prepared single crystal plates of BTO from molten salts and then selectively deposited BiOI on

the positive edges of BTO where there is positive polarization, and thus, a heterojunction has been

achieved where electrons are transferred and injected into BiOI. Because band bending is

generated at the interface between BTO and BiOI, the charge could be spatially separated.

Significant enhancement of photodegradation has been obtained by this method compared with the

conventional loading method. The complete photodegradation of RhB by the BTO-BI-3 sample

requires 12 min under visible light, whereas phenol photodegradation is 47% in 60 min.

6.5 References

1. Al-keisy, A., L. Ren, D. Cui, Z. Xu, X. Xu, X. Su, W. Hao, S.X. Dou, and Y. Du, A ferroelectric

photocatalyst Ag10Si4O13 with visible-light photooxidation properties. Journal of Materials

Chemistry A, 2016. 4(28): p. 10992-10999.

2. Al-Keisy, A., L. Ren, T. Zheng, X. Xu, M. Higgins, W. Hao, and Y. Du, Enhancement of charge

separation in ferroelectric heterogeneous photocatalyst Bi4(SiO4)3/Bi2SiO5 nanostructures.

Dalton Transactions, 2017. 46(44): p. 15582-15588.

3. Hongwei, H., T. Shuchen, Z. Chao, Z. Tierui, R.A. H., and Z. Yihe, Macroscopic Polarization

Enhancement Promoting Photo‐ and Piezoelectric‐Induced Charge Separation and

Molecular Oxygen Activation. Angewandte Chemie International Edition, 2017. 56(39): p.

11860-11864.

4. Zhang, R., Y. Dai, Z. Lou, Z. Li, Z. Wang, Y. Yang, X. Qin, X. Zhang, and B. Huang, Layered

photocatalyst Bi2O2[BO2(OH)] nanosheets with internal polar field enhanced photocatalytic

activity

CrystEngComm, 2014. 16(23): p. 4931-4934.

5. Li, L., P.A. Salvador, and G.S. Rohrer, Photocatalysts with internal electric fields. Nanoscale,

2014. 6(1): p. 24-42.

6. Reza Gholipour, M., C.-T. Dinh, F. Beland, and T.-O. Do, Nanocomposite heterojunctions as sunlight-driven photocatalysts for hydrogen production from water splitting. Nanoscale, 2015.

7(18): p. 8187-8208.

7. T., Y.H., C. T., C.S. G., O.Y. S., and C. S.‐W., Mechanism of the Switchable Photovoltaic

Effect in Ferroelectric BiFeO3. Advanced Materials, 2011. 23(30): p. 3403-3407.

8. Hui, L., C. Jun, R. Yang, Z. Linxing, P. Zhao, F. Longlong, and X. Xianran, Large Photovoltage

and Controllable Photovoltaic Effect in PbTiO3‐Bi(Ni2/3+xNb1/3–x)O3–δ Ferroelectrics.

Advanced Electronic Materials, 2015. 1(4): p. 1400051.

9. Zhen, C., J.C. Yu, G. Liu, and H.-M. Cheng, Selective deposition of redox co-catalyst(s) to

improve the photocatalytic activity of single-domain ferroelectric PbTiO3 nanoplates. Chemical

Communications, 2014. 50(72): p. 10416-10419.

10. Li, S., J. Zhang, B.-P. Zhang, W. Huang, C. Harnagea, R. Nechache, L. Zhu, S. Zhang, Y.-H.

Lin, L. Ni, Y.-H. Sang, H. Liu, and F. Rosei, Manipulation of charge transfer in vertically

aligned epitaxial ferroelectric KNbO3 nanowire array photoelectrodes. Nano Energy, 2017. 35:

p. 92-100. 11. Yang, W., Y. Yu, M.B. Starr, X. Yin, Z. Li, A. Kvit, S. Wang, P. Zhao, and X. Wang,

Ferroelectric Polarization-Enhanced Photoelectrochemical Water Splitting in TiO2–BaTiO3

Core–Shell Nanowire Photoanodes. Nano Letters, 2015. 15(11): p. 7574-7580.

Page 122: P-block-based Ferroelectric-photocatalyst compounds

109

12. Lee, H., H.-y. Joo, C. Yoon, J. Lee, H. Lee, J. Choi, B. Park, and T. Choi, Ferroelectric

BiFeO3/TiO2 nanotube heterostructures for enhanced photoelectrochemical performance.

Current Applied Physics, 2017. 17(5): p. 679-683.

13. Xie, J., C. Guo, P. Yang, X. Wang, D. Liu, and C.M. Li, Bi-functional ferroelectric BiFeO3

passivated BiVO4 photoanode for efficient and stable solar water oxidation. Nano Energy, 2017.

31: p. 28-36.

14. Gupta, S., M. Tomar, and V. Gupta, Ferroelectric photovoltaic response to structural

transformations in doped BiFeO3 derivative thin films. Materials & Design, 2016. 105: p. 296-

300.

15. Wei, J., Y. Kui, and L.Y. C., Bulk Photovoltaic Effect at Visible Wavelength in Epitaxial Ferroelectric BiFeO3 Thin Films. Advanced Materials, 2010. 22(15): p. 1763-1766.

16. Noguchi, Y., T. Goto, M. Miyayama, A. Hoshikawa, and T. Kamiyama, Ferroelectric distortion

and electronic structure in Bi4Ti3O12. Journal of Electroceramics, 2008. 21(1): p. 49-54.

17. Zhang, Y.G., H.W. Zheng, J.X. Zhang, G.L. Yuan, W.X. Gao, Y.Z. Gu, C.L. Diao, Y.F. Liu, and

W.F. Zhang, Photovoltaic effects in Bi4Ti3O12 thin film prepared by a sol–gel method. Materials

Letters, 2014. 125: p. 25-27.

18. He, H., J. Yin, Y. Li, Y. Zhang, H. Qiu, J. Xu, T. Xu, and C. Wang, Size controllable synthesis

of single-crystal ferroelectric Bi4Ti3O12 nanosheet dominated with {001} facets toward enhanced

visible-light-driven photocatalytic activities. Applied Catalysis B: Environmental, 2014.

156(Supplement C): p. 35-43.

19. Zhao, W., Z. Jia, E. Lei, L. Wang, Z. Li, and Y. Dai, Photocatalytic degradation efficacy of Bi4Ti3O12 micro-scale platelets over methylene blue under visible light. Journal of Physics and

Chemistry of Solids, 2013. 74(11): p. 1604-1607.

20. Liu, Y., G. Zhu, J. Gao, M. Hojamberdiev, R. Zhu, X. Wei, Q. Guo, and P. Liu, Enhanced

photocatalytic activity of Bi4Ti3O12 nanosheets by Fe3+-doping and the addition of Au

nanoparticles: Photodegradation of Phenol and bisphenol A. Applied Catalysis B:

Environmental, 2017. 200(Supplement C): p. 72-82.

21. Azodi, M., C. Harnagea, V. Buscaglia, M.T. Buscaglia, P. Nanni, F. Rosei, and A. Pignolet,

Ferroelectric switching in Bi4Ti3O12 nanorods. IEEE Transactions on Ultrasonics, Ferroelectrics,

and Frequency Control, 2012. 59(9): p. 1903-1911.

22. He, H., J. Yin, Y. Li, Y. Zhang, H. Qiu, J. Xu, T. Xu, and C. Wang, Size controllable synthesis

of single-crystal ferroelectric Bi4Ti3O12 nanosheet dominated with {001} facets toward enhanced visible-light-driven photocatalytic activities. Applied Catalysis B: Environmental, 2014. 156-

157: p. 35-43.

23. Cui, Y., J. Briscoe, and S. Dunn, Effect of Ferroelectricity on Solar-Light-Driven Photocatalytic

Activity of BaTiO3—Influence on the Carrier Separation and Stern Layer Formation. Chemistry

of Materials, 2013. 25(21): p. 4215-4223.

24. Zhao, M.H., D.A. Bonnell, and J.M. Vohs, Effect of ferroelectric polarization on the adsorption

and reaction of ethanol on BaTiO3. Surface Science, 2008. 602(17): p. 2849-2855.

25. Zhang, Z., P. Sharma, C.N. Borca, P.A. Dowben, and A. Gruverman, Polarization-specific

adsorption of organic molecules on ferroelectric LiNbO3 surfaces. Applied Physics Letters,

2010. 97(24): p. 243702.

26. Al-Attafi, K., A. Nattestad, Y. Yamauchi, S.X. Dou, and J.H. Kim, Aggregated mesoporous

nanoparticles for high surface area light scattering layer TiO2 photoanodes in Dye-sensitized

Solar Cells. Scientific Reports, 2017. 7(1): p. 10341.

Page 123: P-block-based Ferroelectric-photocatalyst compounds

110

Page 124: P-block-based Ferroelectric-photocatalyst compounds

111

Chapter 7

Conclusions and Recombination for Future Work

7.1 Conclusions

This project included p-block photocatalyst-based internal electric field as shown in research

works which are chapters four, five and six which were organized in three objectives: The first

objective (chapter four) was investigated such as Ag10Si4O13, by materials design and band

engineering. Beside include internal electric generated from distortion SiO4 chains inside crystal

structure, Ag10Si4O13 possessed unique d10 and sp/p configurations in its electronic structure. This

electronic structure promotes the generation, separation, and mobility of photo-induced charge

carriers under visible-light illumination, which has been verified experimentally and theoretically.

The experimentally results are consistent with calculation results that were demonstrated powerful

oxidation potential toward organic dyes and phenol under visible light. This unique electronic

structure and internal electric field originate from the p-block electronic configuration and

distorted SiO4 chains in Ag10Si4O13.

The second objective, photocatalytic activity of ferroelectric materials is highly influenced by the

main direction of charge separation, originating from spontaneous polarization. In this work,

unique bismuth silicate based zerodimensional (0D)/two-dimensional (2D) heterogeneous

nanostructures were successfully constructed. In contrast to either individual pristine phase, this

heterogeneous structure exhibited much enhanced photocatalytic activity towards the degradation

of Rhodamine B and phenol. The synergistic effects of high polarization in 2D ferroelectric

Bi2SiO5 nanosheets and the band bending at the 0D–2D interface of the heterostructures have been

proved to accelerate the photoinduced charge separation and the movement of separated carriers to

the interface, which further improves the photodegradation performance. This work provides a

novel strategy for adjusting the photoinduced carrier transfer route in the ferroelectric materials

and designing novel photocatalysts with ultrafast charge separation and large active surface area.

The third objective, ferroelectric-photocatalyst/photocatalyst heterojunctions have much attractive

toward address challenging of photocatalytic performance, beside charge carrier separation by

internal electric field, charge transfer has been further enhanced by heterojunction interface. In this

work, the polarization-adsorption interaction which exists in ferroelectric materials was employed

for successful deposition of BiOI on a specific surface of Bi4Ti3O12 in the dark at room

temperature, where the positive polarized region of polarization is found. The crystal structure,

morphology and composition of samples were confirmed by X-ray diffraction, field emission

scanning electron microscopy and X-ray photoelectron spectroscopy respectively. The result

shows higher activity by heterojunction, the reason behind enhancement activity was confirmed by

firstly, the band structure contributed to photoelectrons transfer from Bi4Ti3O12 to BiOI, secondly,

increase active site area of positive polar of Bi4Ti3O12, thirdly, eliminate the screening layer which

contribute impedance layer for charge transfer, fourthly, increase visible absorption

Page 125: P-block-based Ferroelectric-photocatalyst compounds

112

In conclusion, the main points of these objective are contribution of internal electric field in

photocatalytic activity and p-block compounds based ferroelectric materials. The internal electric

field which provides by ferroelectric materials due to spontaneous polarization has provided

driving force to enhance charge separation whereas p-block compounds are not only possess

suitable band structure toward high oxidation potential but also possess high mobility charge

charrier.

7.2 Recombination for Future Work

Although this thesis covers three compounds related to p-block based internal electric field but

there are remains some work that would be benfit in future. The following some of these future

works:

➢ Evaluate oxygen and hydrogen evolution for each component. The capability of redox potential

for either water splitting or degradation organic pollutants strictly depends on the positions of

CB and VB, The CB of Ag10Si4O13 has shown lower negatively than reduction potential of

H+/H2, therefore, the hydrogen production by Ag10Si4O13 is excluded, whereas BTO and BSO-

HNS have shown higher negatively than reduction potential of H+/H2, therefore, it is expected to

produce H2. The mineralization of dyes should be investigated by total organic carbon analysis

removal measurement because decolorization of dye is not completely photodegradation.

➢ DFT calculation for a ferroelectric surface with interfacing with a semiconductor in certain face

was hard to calculation with software, the exposed face with existence polarization in crystal

structure interface with expose face for a material is complex to calculate.

Page 126: P-block-based Ferroelectric-photocatalyst compounds

113

Appendices

Appendix 1

List of publications

1. Al-Keisy, A., L. Ren, T. Zheng, X. Xu, M. Higgins, W. Hao, and Y. Du, Enhancement of

charge separation in ferroelectric heterogeneous photocatalyst Bi4(SiO4)3/Bi2SiO5

nanostructures. Dalton Transactions, 2017. 46(44): p. 15582-15588.

2. Al-keisy, A., L. Ren, D. Cui, Z. Xu, X. Xu, X. Su, W. Hao, S.X. Dou, and Y. Du, A

ferroelectric photocatalyst Ag10Si4O13 with visible-light photooxidation properties. Journal

of Materials Chemistry A, 2016. 4(28): p. 10992-10999.

3. Al-keisy, A., et al. Selective deposition of BiOI on the positive polar surface of Bi4Ti3O12

by the polarization-adsorption effect and evaluation of the photocatalytic performance.

Manuscript submitted to applied catalyst b.

Conferences participating

1. 2018 International Symposium on Advanced Materials & Sustainable Technologies at QUT

in Queensland

2. The 2nd International Symposium on Renewable Energy Technologies, Nov. 30 - Dec. 04,

at UTS in Sydney