You are on page 1of 641

Subramanian 

Balakumar
Valérie Keller
M.V. Shankar  Editors

Nanostructured
Materials
for Environmental
Applications
Nanostructured Materials for Environmental
Applications
Subramanian Balakumar  •  Valérie Keller
M. V. Shankar
Editors

Nanostructured Materials for


Environmental Applications
Editors
Subramanian Balakumar Valérie Keller
NCNSNT ICPEES
University of Madras University of Strasbourg
Chennai, India Strasbourg Cedex 2, France

M. V. Shankar
Department of Materials Science
and Nanotech
Yogi Vemana University
Vemanapuram, Andhra Pradesh, India

ISBN 978-3-030-72075-9    ISBN 978-3-030-72076-6 (eBook)


https://doi.org/10.1007/978-3-030-72076-6

© Springer Nature Switzerland AG 2021


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Dedicated to Environmental Scientists
Foreword

Accelerated industrialization and urbanization in the last two centuries have caused
colossal devastation of the pristine environment which we were blessed with.
Attempt to re-establish and recover our habitat is a primal obligation of every indi-
vidual who is a part of this biosphere. Above all, the responsibility is more with the
research fraternity who can transform and revive the environment with the enor-
mous scientific advances available.
The field of nanomaterials has proven to exhibit refreshingly novel opportunities
that can maneuver the catastrophic challenges that the environment is being sub-
jected to. The unique scale which nanomaterials offer has rendered it the most
sought-after approach in creating environmental sustainability. Pathbreaking
advances in nanomaterial synthesis and nanostructure formulation have opened up
newer and promising possibilities to re-create the lost glory.
This book entitled Emerging Nanostructured Materials for Environmental
Application edited by Prof. S. Balakumar and team offers an in-depth diverse range
of emerging nanostructured materials that have been validated for environmental
applications. Key attributes of this book are that the content is exquisite and com-
prehensible. It surpasses similar books of this kind as it is replete with exhaustive
coverage on electrocatalysts, visible light/UV-based photocatalysts, magneto-­
photocatalysis, plasmonic nanostructured catalysts, electrocatalysts, thermo-­
catalysts, organo-catalysts, adsorption of nanostructured materials, chemical and
surface kinetics, degradation of oil and water pollutants, and electromagnetic pol-
lutants and shielding aspects.
This book will serve as an indispensable source of reference to graduate/post-
graduate students and researchers. In addition to placing emphasis on challenges
related to latest advancements, it also elaborates on the trend for further research
and future perspective that is needed for a multidisciplinary approach in environ-
mental science with nanoscale multifunctional materials.

University of Madras S. Gowari


Chennai, Tamil Nadu, India

vii
Acknowledgments

First of all, we place our heartfelt thanks to Almighty, who has been kind enough to
bless us with good health whilst working on this book and for making this task a
great success.
We would like to express our sincere appreciation to Prathik Roy, Group
Product Manager, Nanoscience and Technology, Database Group, Springer Nature,
Jersey City, NJ 07302, who encouraged us to initiate this project and also intro-
duced us to the right person, Dr. Anita Lekhwani, to evaluate our proposal. We are
also immensely grateful to Dr. Anita, who supported our proposal from the begin-
ning and took all the initiative to get the approval of this project. We owe an enor-
mous debt of gratitude to all the authors who contributed to research articles for the
success of the physical creation and completion of this volume.
Dr. S. Balakumar gratefully acknowledges supports received from the University
of Madras. He would also like to extend his warmest thanks to research team mem-
bers of the National Centre for Nanoscience and Nanotechnology for their constant
support and enthusiasm which helped him to complete the task on time.
Dr. M.V. Shankar would like to express his special gratitude and thanks to Prof.
Munagala Surya Kalavathi, Vice Chancellor of Yogi Vemana University, Andhra
Pradesh, India, for her valuable suggestions and constant inspiration. He is also
thankful to alumni and present members of Nanocatalysis and Solar Fuels Research
group for their dedicated research work and contributions to this book.
Dr. Keller acknowledges financial support from ICPEES, Institut de Chimie et
des Procédés pour l’Energie, l ‘Environnement et la Santé, UMR 7515, CNRS
Université de Strasbourg, for this book project.

ix
Contents

1 Nanostructures in Photocatalysis: Opportunities and Challenges


for Environmental Applications��������������������������������������������������������������    1
Y. V. Divyasri, Y. N. Teja, V. Nava Koteswara Rao,
N. C. Gangi Reddy, Sakar Mohan, M. Mamatha Kumari,
and M. V. Shankar
2 Nanostructured Heterojunction (1D-0D and 2D-0D)
Photocatalysts for Environmental Remediation������������������������������������   35
Lakshmana Reddy Nagappagari, Kiyoung Lee, Ajay Rakesh,
Subramanian Balakumar, and M. V. Shankar
3 Hierarchical Nanostructures for Photocatalytic Applications ������������   67
R. Ajay Rakkesh, Durgalakshmi Dhinasekaran, M. V. Shankar,
and S. Balakumar
4 Nanocomposite Photocatalysts for the Degradation of
Contaminants of Emerging Concerns����������������������������������������������������   89
Rokesh Karuppannan, Sakar Mohan, and Trong-On Do
5 Sunlight-Mediated Plasmonic Photocatalysis: Mechanism
and Material Prospects����������������������������������������������������������������������������  119
Durgalakshmi Dhinasekaran, M. R. Ashwin Kishore,
and Mohanraj Jagannathan
6 Photocatalytic Efficiency of Bi-Based Aurivillius Compounds:
Critical Review and Discernment of the Factors Involved������������������  143
Manjunath Shetty, Murthy Muniyappa, M. Navya Rani,
Vinay Gangaraju, Prasanna D. Shivaramu, and Dinesh Rangappa
7 Intrinsically Conducting Polymer Nanocomposites in Shielding
of Electromagnetic Pollution������������������������������������������������������������������  173
Suneel Kumar Srivastava

xi
xii Contents

8 Nanostructuring of Hybrid Materials Using Wrapping Approach


to Enhance the Efficiency of Visible Light-Responsive
Semiconductor Photocatalyst������������������������������������������������������������������  223
V. Vinesh, A. R. Mahammed Shaheer, and B. Neppolian
9 Metal–Organic Frameworks (MOFs) with Hierarchical
Structures for Visible Light Photocatalysis ������������������������������������������  239
P. Karthik and B. Neppolian
10 Soil Remediation by Zero-Valent Iron Nanoparticles
for Organic Pollutant Elimination����������������������������������������������������������  253
Marco Stoller, Luca Di Palma, and Giorgio Vilardi
11 Black TiO2: An Emerging Photocatalyst and Its Applications������������  273
P. Anil Kumar Reddy, P. Venkata Laxma Reddy,
and S. V. Prabhakar Vattikuti
12 Nanomaterials for Photocatalytic Decomposition of Endocrine
Disruptors in Water ��������������������������������������������������������������������������������  305
Ajay Kumar, Vishal Sharma, Ashish Kumar, and Venkata Krishnan
13 Carbonaceous Nanomaterials for Environmental Remediation����������  327
Natarajan Sasirekha and Yu-Wen Chen
14 Magnetically Recyclable Photocatalysts for Degradation
of Organic Pollutants in Aquatic Environment������������������������������������  371
Ashutosh Kumar and Sushil Kumar Kansal
15 Titanate Nanostructures as Potential Adsorbents for
Defluoridation of Water��������������������������������������������������������������������������  389
C. Prathibha, Anjana Biswas, and M. V. Shankar
16 Photocatalytic Water Pollutant Treatment: Fundamental,
Analysis and Benchmarking ������������������������������������������������������������������  407
Katherine Rebecca Davies, Ben Jones, Chiaki Terashima,
Akira Fujishima, and Sudhagar Pitchaimuthu
17 Graphene-Based Photocatalytic Materials: An Overview ������������������  439
Alex T. Kuvarega, Rengaraj Selvaraj, and Bhekie B. Mamba
18 Recent Advances in Nanostructured Materials for Detoxification
of Cr(VI) to Cr(III) for Environmental Remediation��������������������������  461
Udayabhanu, S. B. Patil, and G. Nagaraju
19 Metal Nitrides and Graphitic Carbon Nitrides as Novel
Photocatalysts for Hydrogen Production and Environmental
Remediation����������������������������������������������������������������������������������������������  491
Sudesh Kumar, Kakarla Raghava Reddy, Ch. Venkata Reddy,
Nagaraj P. Shetti, Veera Sadhu, M. V. Shankar,
Vasu Govardhana Reddy, A. V. Raghu, and Tejraj M. Aminabhavi
Contents xiii

20 Highly Functionalized Nanostructured Titanium


Oxide-Based Photocatalysts for Direct Photocatalytic
Decomposition of NOx/VOCs�����������������������������������������������������������������  527
Katchala Nanaji, Manavalan Vijayakumar,
Ammaiyappan Bharathi Sankar, and Mani Karthik
21 Bandgap Engineering as a Potential Tool for Quantum
Efficiency Enhancement��������������������������������������������������������������������������  551
Reddy Kunda Siri Kiran Janardhana, Raju Kumar,
Tata Narsinga Rao, and Srinivasan Anandan
22 Nanostructure Material-Based Sensors for Environmental
Applications����������������������������������������������������������������������������������������������  571
Vinutha Srikanth, Mahesh Shastri, M. Sindhu Sree, M. Navya Rani,
Prasanna D. Shivaramu, and Dinesh Rangappa
23 Nanostructured MoS2 as Non-noble Metal-­Based Cocatalyst
for Photocatalytic Applications��������������������������������������������������������������  597
Murthy Muniyappa, Manjunath Shetty, Mahesh Shastri,
S. Jagadeesh Babu, M. Navya Rani, Prasanna D. Shivaramu,
and Dinesh Rangappa

Index������������������������������������������������������������������������������������������������������������������  611
Contributors

Tejraj M. Aminabhavi  Sonia College of Pharmacy, Dharwad, Karnataka, India


Srinivasan Anandan  Centre for Nano Materials, International Advanced Research
Centre for Powder Metallurgy and New Materials, Hyderabad, Telangana, India
Subramanian Balakumar  National Centre for Nanoscience and Nanotechnology
(NCNN), University of Madras, Chennai, Tamil Nadu, India
Anjana Biswas  Department of Physics, Sri Sathya Sai Institute of Higher Learning,
Anantapur Campus, Anantapur, Andhra Pradesh, India
Yu-Wen  Chen  Department of Chemical and Materials Engineering, National
Central University, Chung-Li, Taiwan
Katherine  Rebecca  Davies  Multi-functional Photocatalyst and Coatings Group,
SPECIFIC, College of Engineering, Swansea University (Bay Campus), Swansea,
Wales, UK
Luca  Di Palma  Department of Chemical Engineering Materials Environment,
Sapienza University of Rome, Rome, Italy
Y. V. Divyasri  Department of Chemistry, Yogi Vemana University, Kadapa, Andhra
Pradesh, India
Trong-On  Do  Department of Chemical Engineering, Laval University, Quebec,
QC, Canada
Durgalakshmi Dhinasekaran  Department of Medical Physics, Anna University,
CEG Campus, Chennai, Tamil Nadu, India
Akira Fujishima  Photocatalysis International Research Center, Tokyo University
of Science, Nodashi, Chiba ken, Japan
Vinay Gangaraju  Department of Applied Sciences, Visvesvaraya Technological
University, Center for Postgraduate Studies, Muddenahalli, Chikkaballapur,
Karnataka, India
S. Jagadeesh Babu  Department of Applied Sciences, Visvesvaraya Technological
University, Center for Postgraduate Studies, Muddenahalli, Chikkaballapur,
Karnataka, India

xv
xvi Contributors

Mohanraj Jagannathan  Department of Medical Physics, Anna University, CEG


Campus, Chennai, Tamil Nadu, India
Reddy Kunda Siri Kiran Janardhana  Centre for Nano Materials, International
Advanced Research Centre for Powder Metallurgy and New Materials, Hyderabad,
Telangana, India
Ben  Jones  Multi-functional Photocatalyst and Coatings Group, SPECIFIC,
College of Engineering, Swansea University (Bay Campus), Swansea, Wales, UK
Sushil  Kumar  Kansal  Dr. S.  S. Bhatnagar University Institute of Chemical
Engineering and Technology, Panjab University, Chandigarh, India
Mani Karthik  Centre for Nanomaterials, International Advanced Research Centre
for Powder Metallurgy and New Materials (ARCI), Balapur, Hyderabad,
Telangana, India
P.  Karthik  SRM Research Institute, SRM Institute of Science and Technology,
Kattankulathur, Chennai, Tamil Nadu, India
Rokesh  Karuppannan  Department of Chemical Engineering, Laval University,
Quebec, QC, Canada
M. R. Ashwin Kishore  Department of Chemical Engineering, University of Seoul,
Seoul, Republic of Korea
Venkata  Krishnan  School of Basic Sciences and Advanced Materials Research
Center, Indian Institute of Technology Mandi, Kamand, Himachal Pradesh, India
Ajay Kumar  School of Basic Sciences and Advanced Materials Research Center,
Indian Institute of Technology Mandi, Kamand, Himachal Pradesh, India
Ashish  Kumar  School of Basic Sciences and Advanced Materials Research
Center, Indian Institute of Technology Mandi, Kamand, Himachal Pradesh, India
Ashutosh  Kumar  School of Energy and Environment, Thapar Institute of
Engineering and Technology, Patiala, Punjab, India
M.  Mamatha  Kumari  Nanocatalysis and Solar Fuels Research Laboratory,
Department of Materials Science & Nanotechnology, Yogi Vemana University,
Kadapa, Andhra Pradesh, India
Raju Kumar  Centre for Nano Materials, International Advanced Research Centre
for Powder Metallurgy and New Materials, Hyderabad, Telangana, India
Sudesh  Kumar  Department of Chemistry, Banasthali Vidyapeeth, Vanasthali,
Rajasthan, India
Alex  T.  Kuvarega  Nanotechnology and Water Sustainability Research Unit,
College of Science, Engineering and Technology, University of South Africa,
Florida Science Campus, Florida, Johannesburg, South Africa
Contributors xvii

Kiyoung  Lee  Department of Energy Chemical Engineering, School of Nano &


Materials Science and Engineering, Kyungpook National University, Sangju,
Republic of Korea
Bhekie  B.  Mamba  Nanotechnology and Water Sustainability Research Unit,
College of Science, Engineering and Technology, University of South Africa,
Florida Science Campus, Florida, Johannesburg, South Africa
Murthy Muniyappa  Department of Applied Sciences, Visvesvaraya Technological
University, Center for Postgraduate Studies, Muddenahalli, Chikkaballapur,
Karnataka, India
Lakshmana Reddy Nagappagari  Department of Energy Chemical Engineering,
School of Nano & Materials Science and Engineering, Kyungpook National
University, Sangju, Republic of Korea
G.  Nagaraju  Energy Materials Research Laboratory, Department of Chemistry,
Siddaganga Institute of Technology, Tumakuru, Karnataka, India
Katchala  Nanaji  Centre for Nanomaterials, International Advanced Research
Centre for Powder Metallurgy and New Materials (ARCI), Balapur, Hyderabad,
Telangana,, India
B.  Neppolian  Energy and Environmental Remediation Lab, SRM Research
Institute, SRM Institute of Science and Technology, Kattankulathur, Chennai, Tamil
Nadu, India
S.  B.  Patil  Energy Materials Research Laboratory, Department of Chemistry,
Siddaganga Institute of Technology, Tumakuru, Karnataka, India
Department of Chemistry, The Oxford College of Science, Bengaluru,
Karnataka, India
Sudhagar  Pitchaimuthu  Multi-Functional Photocatalyst and Coatings Group,
SPECIFIC, College of Engineering, Swansea University (Bay Campus), Swansea,
Wales, UK
C. Prathibha  Department of Physics, Sri Sathya Sai Institute of Higher Learning,
Anantapur Campus, Anantapur, Andhra Pradesh, India
A. V. Raghu  Department of Chemistry, Faculty of Engineering and Technology,
Jain (Deemed-to-be University), Bangalore, Karnataka, India
Ajay  Rakesh  National Centre for Nanoscience and Nanotechnology (NCNN),
University of Madras, Chennai, Tamil Nadu, India
R.  Ajay  Rakkesh  National Centre for Nanoscience and Nanotechnology,
University of Madras, Chennai, Tamil Nadu, India
Department of Physics and Nanotechnology, SRM Institute of Science and
Technology, Kattankulathur, Tamil Nadu, India
xviii Contributors

Dinesh Rangappa  Department of Applied Sciences, Visvesvaraya Technological


University, Center for Postgraduate Studies, Muddenahalli, Chikkaballapur,
Karnataka, India
M. Navya Rani  School of Basic and Applied Sciences, Dayanand Sagar University,
Bengaluru, Karnataka, India
Tata Narsinga Rao  Centre for Nano Materials, International Advanced Research
Centre for Powder Metallurgy and New Materials, Hyderabad, Telangana, India
V.  Nava  Koteswara  Rao  Nanocatalysis and Solar Fuels Research Laboratory,
Department of Materials Science & Nanotechnology, Yogi Vemana University,
Kadapa, Andhra Pradesh, India
Ch.  Venkata  Reddy  School of Mechanical Engineering, Yeungnam University,
Gyeongsan, South Korea
Kakarla  Raghava  Reddy  School of Chemical and Biomolecular Engineering,
The University of Sydney, Sydney, NSW, Australia
N. C. Gangi Reddy  Department of Chemistry, Yogi Vemana University, Kadapa,
Andhra Pradesh, India
P.  Anil  Kumar  Reddy  School of Mechanical and Nuclear Engineering, Ulsan
National Institute of Science and Technology (UNIST), Ulsan, South Korea
P. Venkata Laxma Reddy  Program in Environmental Science and Engineering,
University of Texas El Paso, El Paso, TX, USA
Vasu  Govardhana  Reddy  Department of Chemistry, Yogi Vemana University,
Kadapa, Andhra Pradesh, India
Veera Sadhu  School of Physical Sciences, Kakatiya Institute of Technology and
Science (KITS), Warangal, Telangana, India
Sakar Mohan  Centre for Nano and Material Sciences, Jain University, Bangalore,
Karnataka, India
Department of Chemical Engineering, Laval University, Quebec, QC, Canada
Ammaiyappan  Bharathi  Sankar  School of Electronics Engineering, Vellore
Institute of Technology (VIT), Chennai Campus, Chennai, Tamil Nadu, India
Natarajan  Sasirekha  CAS in Crystallography and Biophysics, University of
Madras, Guindy Campus, Chennai, Tamil Nadu, India
Rengaraj  Selvaraj  Chemistry Department, Sultan Qaboos University, Muscat,
Sultanate of Oman
A. R. Mahammed Shaheer  Energy and Environmental Remediation Lab, SRM
Research Institute, SRM Institute of Science and Technology, Kattankulathur,
Chennai, Tamil Nadu, India
Contributors xix

M. V. Shankar  Nanocatalysis and Solar Fuels Research Laboratory, Department


of Materials Science & Nanotechnology, Yogi Vemana University, Kadapa, Andhra
Pradesh, India
Vishal  Sharma  School of Basic Sciences and Advanced Materials Research
Center, Indian Institute of Technology Mandi, Kamand, Himachal Pradesh, India
Mahesh  Shastri  Department of Applied Sciences and Visvesvaraya Center for
Nanoscience and Technology, Centre Post Graduation Studies, Visvesvaraya
Technological University, Muddenahalli Campus, Chikkaballapura, India
Nagaraj  P.  Shetti  Department of Chemistry, K.  L. E.  Institute of Technology,
Hubli, Karnataka, India
Visvesvaraya Technological University, Belgaum, Karnataka, India
Manjunath Shetty  Department of Applied Sciences, Visvesvaraya Technological
University, Center for Postgraduate Studies, Muddenahalli, Chikkaballapur,
Karnataka, India
Prasanna  D.  Shivaramu  Department of Applied Sciences, Visvesvaraya
Technological University, Center for Postgraduate Studies, Muddenahalli,
Chikkaballapura, India
M. Sindhu  Sree  Department of Applied Sciences and Visvesvaraya Center for
Nanoscience and Technology, Centre Post Graduation Studies, Visvesvaraya
Technological University, Muddenahalli Campus, Chikkaballapura, India
Vinutha Srikanth  Department of Electrical and Electronics Engineering, KSSEM,
Bengaluru, Karnataka, India
Suneel  Kumar  Srivastava  Department of Chemistry, Indian Institute of
Technology, Kharagpur, West Bengal, India
Marco  Stoller  Department of Chemical Engineering Materials Environment,
Sapienza University of Rome, Rome, Italy
Y.  N.  Teja  Centre for Nano and Material Sciences, Jain University, Bangalore,
Karnataka, India
Chiaki Terashima  Photocatalysis International Research Center, Tokyo University
of Science, Nodashi, Chiba ken, Japan
Udayabhanu  Energy Materials Research Laboratory, Department of Chemistry,
Siddaganga Institute of Technology, Tumakuru, Karnataka, India
Center for Research and Innovations, BGSIT, Adichunchanagiri University,
B. G. Nagara, Mandya, Karnataka, India
S.  V.  Prabhakar  Vattikuti  School of Mechanical Engineering, Yeungnam
University, Gyeongsan, South Korea
xx Contributors

Manavalan  Vijayakumar  Centre for Nanomaterials, International Advanced


Research Centre for Powder Metallurgy and New Materials (ARCI), Balapur,
Hyderabad, India
Global Innovative Centre for Advanced Nanomaterials (GICAN), Collage of
Science, Engineering and Environment, The University of Newcastle, NSW,
Callaghan, Australia
Giorgio  Vilardi  Department of Chemical Engineering Materials Environment,
Sapienza University of Rome, Rome, Italy
V. Vinesh  Energy and Environmental Remediation Lab, SRM Research Institute,
SRM Institute of Science and Technology, Kattankulathur, Chennai, Tamil
Nadu, India
About the Editors

Subramanian Balakumar, FRSC, FASCh.  is a Professor and the Director of the


National Centre for Nanoscience and Nanotechnology, University of Madras. He is
a distinguished academician and a researcher par excellence in the field of nanosci-
ence. He received his Ph.D. in the Faculty of Sciences, from the prestigious Anna
University, India, in 1996 and further escalated his research competence as a post-
doctoral fellow with Prof. J.B.  Xu, Department of Electronic Engineering, the
Chinese University of Hong Kong, Hong Kong, from 1995 to 1997, and subse-
quently worked as NSTB postdoctoral fellow with Prof. Zeng Hua Chen, Department
of Chemical Engineering, National University of Singapore, from 1997 to 1999. His
remarkable stint in industry began as a Senior Research Engineer in Chartered
Semiconductor Manufacturing Ltd., Singapore (now GLOBALFOUNDRIES,
Singapore), till 2002 and consecutively worked as Senior Scientist at A-Star Institute
of Microelectronics till 2008. In addition to his presence in industry, he remained
deeply connected with academia as an Adjunct Associate Professor, Department of
Mechanical Engineering, National University of Singapore, from 2004 to 2008.
With intense urge in academics, he joined the highly renowned NCNSNT, University
of Madras, during November 2008. Predictably, he established an unsurpassable
academic footprint with more than 250 papers in peer-reviewed journals and also
has more than 8 US patents to his credit. He has published several book chapters and
edited many books. His versatile administrative capabilities were emblazoned by
organizing 40 international and national conferences, workshops, seminars, short-
term courses, and seminars during the past 12 years. His formidable research con-
duct fetched him the role of Associate Editor of Chemical Papers Journal (Springer),
and he serves as an editorial board member for various journals. He has received
more than 20 funded projects from various funding agencies. As an epitome of his
credits, he was recently bestowed with Tamil State Best Scientist Award 2018
(TANSA 2018) and received the Senior Scientist Award from TNHE, Science City,
during 2019. He was also conferred fellow of Royal Society of Chemistry. His
research work includes nanostructured materials for energy, environment, and bio-
medical applications.

xxi
xxii About the Editors

Valérie  Keller  is a senior scientist (Research Director) at ICPEES, Institute of


Chemistry and Processes for Energy, Environment and Health in Strasbourg
(France). She received her Ph.D. degree in Chemistry and Catalysis from the Louis
Pasteur University of Strasbourg in 1993. In 1996, she returned to Strasbourg and
was appointed as researcher in CNRS, where she was promoted as Research Director
in 2012. She is now responsible for the team “Photocatalysis, Photoconversion and
Green Chemical Processes.” Her main research activities concern photocatalysis for
environmental, energy, and health applications and the synthesis and characteriza-
tion of nanomaterials for photoconversion purposes. She is the author of over 130
publications in peer-reviewed journals and more than 100 oral communications in
international conferences and symposium. She is also the author of 15 patents. In
2013, she was awarded the First Prize of the Strategic Reflection (awarded by the
French Prime Minister, Manuel Valls). Since 2016, she has been co-director of the
French Solar Fuels (GDR) Network.

M. V. Shankar  is a Professor of Materials Science and Nanotechnology in Yogi


Vemana University, Kadapa, Andhra Pradesh, India. He is leading Nanocatalysis
and Solar Fuels research group in YV University. Renowned for his outstanding
research work in the field of photocatalysis for hydrogen fuel production and multi-
functional application of nanomaterials, he has authored/coauthored 122 publica-
tions and 6 patents with an h-index of 36, i-10 index of 65, and average impact
factor of 4.69 with 6124+ citations. Highlight of his research work is the use of
earth-abundant materials, simple concepts, and scalable experimental methods in
the preparation of photocatalysts. He has proved that crude glycerol- and sulfide-­
containing wastewater can be used as a source for large-scale hydrogen generation
under sunlight. He has chaired many national and international conferences, semi-
nars, symposia, workshops, and DST-INSPIRE programs. He has delivered several
keynotes and invited lectures. Dr. Shankar has very good R&D collaborators in the
area of energy, environment, and healthcare applications. He has rich research expe-
rience in countries, viz., France, Japan, and Taiwan, and is a recipient of several
prizes, awards, and fellowships. He is a recipient of several awards, including
Fellow of Indian Chemical Society, Charted Chemist, and Fellow of Royal Society
of Chemistry (FRSC), London. Dr. Shankar’s research work has been highlighted in
national newspapers and TV channels, earning the mark of exceptional work.
Chapter 1
Nanostructures in Photocatalysis:
Opportunities and Challenges
for Environmental Applications

Y. V. Divyasri, Y. N. Teja, V. Nava Koteswara Rao, N. C. Gangi Reddy,


Sakar Mohan, M. Mamatha Kumari, and M. V. Shankar

1.1  Introduction

The environment, which is essentially the air, water, and soil, is largely polluted due
to the increased population and industrialization. These pollutants are mostly
anthropogenic, and they generally include (i) the toxic-organic materials such as
dyes, aromatic, and aliphatic molecules; (ii) agricultural wastages such as the pesti-
cides, insecticides, and herbicides; (iii) plastics; (iv) pharmaceutical products and
byproducts; (v) inorganic materials such as heavy metals; (vi) toxic gases such as
CO, SOx, and NOx; and (vii) microorganisms such as bacteria, viruses, and fungi
[1–3]. Release of these pollutants into the environment from various sources causes
much adverse effects to the ecology, and it will make permanent damages and even
more worse adverse effects if these pollutants are accumulated into the environ-
ment. Therefore, it is an urgent requirement to address such issues toward destruct-
ing and converting these pollutants into nontoxic. Considering current scenario of
energy consumption, the world also requires energy- and cost-effective techniques
to address the issues in the environmental remediation. In this aspect, photocatalysis
is one of the reliable energy- and cost-effective and versatile techniques, which can
almost degrade/convert into nontoxic/kill all of the abovementioned various catego-
ries of pollutants in the environment [4, 5].

Y. V. Divyasri · N. C. Gangi Reddy


Department of Chemistry, Yogi Vemana University, Kadapa, Andhra Pradesh, India
e-mail: ncgreddy@yogivemanauniversity.ac.in
Y. N. Teja · S. Mohan
Centre for Nano and Material Sciences, Jain University, Bangalore, Karnataka, India
e-mail: m.sakar@jainuniversity.ac.in
V. Nava Koteswara Rao · M. Mamatha Kumari · M. V. Shankar (*)
Nanocatalysis and Solar Fuels Research Laboratory, Department of Materials Science
& Nanotechnology, Yogi Vemana University, Kadapa, Andhra Pradesh, India
e-mail: shankar@yogivemanauniversity.ac.in

© Springer Nature Switzerland AG 2021 1


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_1
2 Y. V. Divyasri et al.

Photocatalysis is the process that leads to the oxidation and reduction reactions,
where it is being done using a photoactive semiconductor (photocatalyst) by activat-
ing it using a suitable light source such as UV and/or visible light. The research in
photocatalysis was started after the discovery of water splitting using the illumi-
nated TiO2 by Fujishima and Honda in 1972 [6]. This profound observation revealed
the concept of photo-induced reduction-oxidation (redox) reactions on the semicon-
ductors. Eventually, it was realized that the redox reactions can be effectively
applied to the environmental remediation applications when the photocatalytic oxi-
dation of CN− and SO−3 was demonstrated on various semiconductors. Later, TiO2
was demonstrated for its photocatalytic ability to degrade the chlorinated com-
pounds [7, 8] and kill microorganisms [9]. Research in photocatalysis has been
advanced in many different ways such as the development of new processes such as
advanced oxidation process (AOP) [10], Fenton’s process [11], photocatalytic
memory effect [12], and new material systems such as heterojunction [13], plas-
monic [14, 15], Z-Scheme [16], and ferroelectric-photocatalysts [17]. The require-
ment of different photocatalytic systems is essentially to produce redox species with
suitable energy to destroy various pollutants. Therefore, the photocatalyst design,
especially at nanoscale, has become important, and as a result there have been many
synthesis methods emerged toward the synthesis of various nanostructures of pho-
tocatalytic systems.
In this chapter, we have described the photocatalytic mechanism toward environ-
mental remediation, various synthesis methods to produce nanostructured photo-
catalysts, and various photocatalytic applications toward environmental remediation
that include the degradation of a range of pollutants such as dyes, pharmaceutical,
pesticides, volatile compounds, plastics; detoxification of various heavy metals and
gases; conversion of greenhouse gas CO2 into hydrocarbon fuels; and killing of
pathogens such as bacteria, fungi, etc. Finally, it concludes with the summary and
future perspective on how the nanostructuring of photocatalysts can improve the
photocatalytic phenomenon and process.

1.2  Mechanics of Photocatalysis in Nanostructures

The mechanism of photocatalysis generally involves the excitation of electron-hole


pairs in a photocatalytic semiconductor under suitable light energy, where they fur-
ther get promoted to the surrounding and produce the radical species to perform the
reduction and oxidation reactions toward the desirable photocatalytic applications.
In this process, the relative band edge position of the photocatalyst with respect to
the redox species is very much important to have the desirable photocatalytic pro-
cess. For instance, the photocatalytic water splitting requires the conduction band
should be more positive and the valence band should be more negative. Similarly,
the specific application needs specific type of radical species that essentially origi-
nates either electrons or holes. For instance, dye degradation process happens
through oxidation reactions, where it requires the participation of holes, while the
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 3

toxic heavy metals can be converted into nontoxic via reduction reactions that
require the electrons. Considering the process of destructing the pollutants, the
required photocatalytic reactions can be depicted as shown in Fig. 1.1, and it can be
described in the following Eqs. (1.1)–(1.9).

PC + hv → e − + h + (1.1)

h + + H 2 O → H + + OHÙ (1.2)

h + + OH − → OHÙ (1.3)

e − + O2 → O2 − (1.4)

2e − + O2 + 2H + → H 2 O 2 (1.5)

2e − + H 2 O2 → OHÙ+OH − (1.6)

R + OHÙ → Degradation products (1.7)

R + h+ → Oxidation degraded products (1.8)

R + e− → Reduction degraded products (1.9)

These abovementioned reactions and mechanisms are about the chemical reac-
tions that occur in the process. However, these reactions can be effectively con-
trolled or enhanced through physical structuring of photocatalysts, where the
potential energy of generated electrons and holes thereby the radical species is often
determined by the energy band edge (VB and CB) positions in the given photocata-
lyst. For instance, the band edge positions can be effectively tuned via reducing the
particle size and changing the morphology of the photocatalysts as shown in
Fig. 1.2(a)–(c), where it shows that when the layer thickness of g-C3N4 is reduced
from its bulk structure to the nanomesh-like structure (a), the relative band edge
positions get shifted (b, c) in g-C3N4 [18].
On the other hand, the band structures can be modified via doping and forming
heterojunction, etc., where the former introduces new energy levels between the VB
and CB, while the latter creates a kind of cascading process/channels to transport
the excited carriers for the required redox reactions [13, 19, 20]. However, the

Fig. 1.1 Photocatalytic
mechanism toward the
environmental remediation
applications
4 Y. V. Divyasri et al.

Fig. 1.2 (a) Exfoliation of bulk g-C3N4 into nanomesh, (b) XPS valence band spectra, and (c) rela-
tive shift in the band edge position in nanomesh of g-C3N4 as compared to its bulk [18]

nanostructure-induced enhancement in the photocatalytic process is relatively


exotic as it leads to geometrics-dependent optical and electronic properties. For
instance, recombination of charge carriers in one-dimensional nanostructures is
relatively enhanced as compared to the spherical nanostructures, which is essen-
tially due to the enhanced delocalization of electrons in the CB of one-­dimensionally
structured photocatalysts.

1.3  Synthesis of Nanostructured Photocatalysts

Synthesis of nanostructures can be broadly divided into two categories such as (i)
top-down and (ii) bottom-up approaches. Most of the photocatalytic nanostructures
have been prepared via bottom-up approach due to the specific reasons such as dop-
ing, composite formation, etc. Moreover, the bottom-up approach-mediated synthe-
ses of nanomaterials pave ways to effectively control the dimension and morphology
of the nanostructures. This section provides a glimpse on some of the selective
synthesis methods (from the recent reports) toward synthesizing various photocata-
lytic nanostructures such as quantum dots, nanospheres, nanotubes, nanocubes,
nanoneedles, nanofibers, nanoporous structures, nanosheets, nanoflowers, and other
anisotropic nanostructures that are successfully used for a variety of photocatalytic
applications.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 5

1.3.1  Sol–Gel Synthesis

Sol-gel process generally involves the transformation of precursor solution into gel.
In the first step of process, the metal precursors are hydrolyzed thereby forming the
metal hydroxide matrix solution. Typically in the second step, the solution is con-
densed where the formation of gel takes place. Thereafter the gel is subjected to the
process of drying or calcination where desired nanostructures are obtained.
Extensive use of this method is done for synthesis of different nanostructures as it
gives the advantage of controlling the morphology such as size and shape at low
temperatures. Flake-like BiVO4 nanostructures were prepared using bismuth nitrate
(Bi (NO3)3·5H2O) and ammonium vanadate (NH4VO3) as initial precursors and
were used under visible light irradiation for the degradation of methylene blue (MB)
[21]. Homogeneous uniformly dispersed spherical−/elliptical-shaped cadmium
(Cd)-doped cerium oxide (CeO2) nanostructures with ferromagnetic nature were
synthesized for the purpose of dye degradation using sol–gel method with cerium
chloride heptahydrate (CeCl3∙7H2O) and cadmium chloride (CdCl2) as starting pre-
cursors with varying doping concentration of 1–3% [22]. The nanostructures
showed an efficiency of 86.42% and 92.53% for the dye degradation of rhodamine
B (RhB) and Congo red (CR) dyes, respectively, within 6-h irradiation under UV–
Vis irradiation. In another example, spindle-like ZnO nanostructures codoped with
neodymium (Nd) and vanadium (V), synthesized by ultrasonic-assisted sol–gel
method, showed excellent efficiency for the photocatalytic degradation of organic
pollutants under visible light irradiation [23]. Pure ZnO and Nd mono-doped ZnO
nanostructures showed perfect spindle-like morphology, whereas the ZnO particles
codoped with 4% Nd and 1% V showed more like a needle-shaped morphology
with reduced diameter size. Using sol–gel method, TiO2–CdO–Ag nanocomposites
were synthesized with an average particle size of 20  nm but with irregular mor-
phologies [24]. The nanocomposites showed almost 92% dye degradation of methy-
lene blue within 75-min irradiation under visible light. Also for the photocatalytic
degradation of organic dyes, Cd2V2O7 nanostructures with an average diameter size
of 10–20 nm were prepared [25]. These results also evidence that the sol-gel process
could be effective to produce nanostructured materials in large scale as well.

1.3.2  Hydro-/Solvo-thermal Synthesis

Hydrothermal synthesis process involves the performance of chemical reactions


among precursors dissolved in the aqueous solvent under autogenous pressure,
where the temperature of the solvent is maintained above the critical point. Solvo-­
thermal process is also similar to hydrothermal except that the solvent used is non-
aqueous. Most of the hydro−/solvo-thermal processes are carried out in a metal
autoclave made up of Teflon or alloy linings. For example, flowerlike ZnO nano-
structures were synthesized using hydrothermal process [26]. In this procedure, two
6 Y. V. Divyasri et al.

10 mL aqueous solutions, one with zinc acetate and another with potassium hydrox-
ide (KOH), were prepared separately. Both solutions were added to double-distilled
water and stirred vigorously together, and the resulting mixed solution was trans-
ferred into a Teflon-lined stainless steel autoclave and was heated for 8 h at a tem-
perature of 120 °C. The resulting precipitate was centrifuged and washed with water
and alcohol for several times to obtain a balanced pH and was finally dried at 60 °C
for 24 h in an oven. It is observed that the obtained nanostructures exhibited flower-
like uniform morphology with an average diameter of 2–4 μm and were used for
photocatalytic dye degradation of methylene blue dye. Core-shell CeO2 nanospheres
were successfully synthesized using template-free hydrothermal method for the
methyl orange dye degradation [27]. The average size of nanospheres synthesized
with 20 mmol of urea was found to be 50–145 nm and showed degradation effi-
ciency as high as 93.49% under UV irradiation of 160  min. Other examples of
hydrothermally synthesized nanostructures are SnO2@ZnO [28] and CuO@ZnO
[29] heterojunctions for H2O evolution and dye degradation, respectively. Using
solvo-thermal method, Rh-In2O3 3D urchin-like and 2D rodlike nanostructures with
an average diameter size of 500–600 nm and 20 nm–1 μm, respectively, were syn-
thesized [30]. Other examples of solvo-thermally synthesized nanostructures
include flowerlike BiVO4 [31] and CdS nanostructures [32], ZnO nanospheres, and
hexagonal disklike ZnO nanostructures [33].

1.3.3  Precipitation Process

Precipitation method is a cost-effective and low-temperature approach used for the


synthesis of photocatalytic nanostructures. Typically, precipitation synthesis method
involves three steps. In the first step, two types of solutions, one with inorganic
metal salt precursor and another one with a chosen surfactant, are prepared sepa-
rately by dissolving in water. In second step, the two solutions are added and stirred
vigorously, and a basic solution such as NaOH is added in a dropwise manner under
stirring and a precipitate is formed. In the next step, the precipitate is collected by
centrifuging and washed with ethanol and water several times to balance the pH
value, and finally the precipitate is dried in an oven followed by calcinations in
some cases. For example, CuO nanosheets were prepared by dissolving copper sul-
fate pentahydrate (CuSO4·5H2O) and cetyltrimethylammonium bromide (CTAB) in
deionized water followed by vigorous stirring and addition of NH3·H2O and NaOH
one after the other in a dropwise manner [34]. After stirring the solution for a certain
amount of time, a blue-colored precipitate was formed which was kept at room
temperature for 2 days to settle down completely, and finally a black-colored pre-
cipitate was obtained. It was washed with ethanol and water several times, centri-
fuged, and finally dried at 80 °C for 6 h. The nanosheets showed an efficiency of
82% in dye degradation under visible light irradiation. Other examples of photo-
catalytic nanostructures synthesized through precipitation approach include CeO2/
Au/Ho cube-like and pyramid-like nansostructures [35], spherical holmium oxide
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 7

(HoO3) nanoparticles [36], neodymium oxide (Nd2O3) bundle-like and spherical


nanostructures [37], and SnO nanosheets [38].

1.3.4  Anodization Process

Anodization process involves development of an oxide layer on the surface of metal


using electrolytic reactions. Even though anodization was employed for synthesis of
various mesoporous nanostructures, TiO2 nanotubes synthesized through anodiza-
tion are most exploited for photocatalysis. In 1999 Zwilling et al. first reported the
synthesis of self-assembled TiO2 nanotubes by anodization [39]. Generally the
anodization setup consists of two electrode system with Ti as anode and inert metal
as cathode (mostly Pt is used), an acidic electrolyte (mostly fluoride ion electrolyte
is used and a DC power supply as shown in Fig. 1.3). The process of nanotube for-
mation mainly involves the simultaneous reactions of oxidation and dissipation.
First electrochemical oxidation of Ti into TiO2 takes place, which forms a compact
oxide layer on the surface of Ti foil, and then due to fluorine ions (F−), small pores
or pits on the compact layer are formed which reduces the resistance toward the
applied current, and finally an equilibrium between the nanoporous oxide layer and
chemical dissipation due to electrical field induction as well as fluorine ion-induced
dissipation leads to the formation of nanotubes. Morphological properties such as
size, shape, and crystalline phase of nanotubes are influenced by various anodiza-
tion parameters such as type of electrolyte and electrodes, pH value of electrolyte,
applied voltage, temperature, and current density.

Fig. 1.3  Schematic representation of anodization setup for synthesis of TiO2 nanotubes [41]
8 Y. V. Divyasri et al.

TiO2 nanotubes were synthesized for degradation of methylene blue under UV


irradiation using Ti foil as anode and stainless steel foil as cathode, respectively, and
an electrolyte mixture of ethylene glycol, deionized water, and NH4F [40].
Nanotubes were synthesized after various voltages of 20–60  V was supplied for
24 h at a temperature of 30 °C. It is observed that with increase in voltage there was
an increase in the diameter size of the nanotubes: 20 V, 80 nm; 40 V, 110 nm; and
60 V, 130 nm. Iron oxide nanostructures for the purpose of water splitting were also
synthesized by anodization using iron rods and Pt foil as electrodes and an electro-
lyte same as in case of TiO2, but here different temperatures have been employed in
order to observe the morphological changes [41]. At the temperature of 25 °C, sta-
ble nanotube structures were observed, but when the temperature was increased
slowly, the morphology shrunk and finally at a temperature of 60 °C granular layer-­
like nanostructures were observed.

1.3.5  Electrospinning

Electrospinning is widely used for the synthesis of nanofibers. It is a simple, cost-­


effective, and quick method to produce nanofibers of diameter size from submicron-­
size to nano-size by applying a high electric DC field to a polymer solution or
melting at room temperature. Electrospun nanofibers carry various advantages such
as higher surface area, high porosity, and small pore size, which are highly favorable
for photocatalysis. The electrospinning setup is shown in Fig. 1.4.
An electrically charged polymer jet is created when high electric DC field is
applied to the polymer solution or melt. When the electric field is further increased,
the polymer jet experiences stability changes where it is stretched before it reaches

Pump

Syringe

Composite Solution Solution Parameters


Viscosity/Concentration
Needle Conductivity
Molecular weight
Taylor Cone
Power Supply Surface Tension
(High voltage) Solution Jet (stability) Solvent Selection

V Solution Jet Electrospinning Parameters


(instability) Voltage Supply
/ Needle Diameter t
whipping Flow Rate
Collector
Nanofibers
Ambient parameters
Temperature
Humidity
Collector

Fig. 1.4  Schematic representation of electrospinning setup with various parameters [42]
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 9

the collector placed just below. Nanofibers are obtained on the collector where the
polymer jet is stiffened into fibers by evaporation of solvent. The diameter size and
the functionality of the nanofibers can be altered by controlling different parameters
mentioned in Fig. 1.4.
CuO/SnO2 hallow nanofibers were developed through electrospinning followed
by annealing for the degradation of methylene blue under visible light irradiation
[43]. The obtained fiber morphology consisted of rough surfaces with average diam-
eter size of 400–600 nm. α-Fe2O3 nanowires with an average diameter of 100 nm
and with pipelike hallow morphology were prepared for photocatalytic dye degra-
dation of rhodamine B and for water splitting through electrospinning assisted by
calcination [44]. The nanowires obtained by calcination at 550 °C showed excellent
photocatalytic efficiency than the samples obtained at 700 °C. Other examples of
photocatalytic nanostructures synthesized by electrospinning method include
ZnFe2O4 nanotubes and nanobelts for photodegradation of rhodamine B [45], lan-
thanum (La)-doped ZnO ceramic nanostructures for degradation of Congo red dye
[46], and Au/ZnO nanostructures for organic pollutant degradation [47].

1.3.6  Pechini Method

Pechini method is considered as a special case of sol–gel method used to obtain


high degree homogeneous end products. In a typical method of Pechini process, a
chelate is formed between the cationic salts or desired precursors that are dissolved
in water and hydroxycarboxylic acid (citric acid is used mostly). Then, addition of
polyalcohol and heating of obtained solution result in the esterification which leads
to the formation of a gel kind of mixture. The final step involves drying of the gel
followed by calcinations after which the desired nanomaterials are obtained.
Gd2CoMnO6 nanostructures were synthesized using Pechini method for the pho-
tocatalytic dye degradation applications [48]. In a typical method, gadolinium
nitrate (GdNO3) and citric acid were dissolved in distilled water followed by the
addition of manganese nitrate hexahydrate Mn(NO3)2·6H2O and cobalt nitrate
(CoNO3)2 under constant stirring. Then the solution was heated followed by the
addition of propylene glycol after which a high viscous gel was obtained. Finally,
the gel was dried in an oven followed by calcination where agglomerated nanopar-
ticles with average size of 25–100 nm were obtained. Nd2Sn2O7 nanostructures with
various morphologies such as less uniform nano-bundles, flake-like nanostructures,
and uniform spherical nanoparticles were synthesized using different stabilizing
agents for photodegradation of methylene orange under UV irradiation [49].
Zirconium nanosheets [50], sphere-like Cu2O/Li3BO3, and CuO/Li3BO3 nanocom-
posites with an average diameter size of 20 nm [51] and urchin-like Dy2CoMnO6
double perovskite nanostructures [52] were also synthesized through Pechini
method for the purpose of dye degradation.
10 Y. V. Divyasri et al.

1.3.7  Other Methods

In the recent years, microemulsion technique has emerged as an interesting way to


synthesize nanostructured photocatalysts owing to its promising advantage of con-
trolling the particle size and morphology. Typically, two immiscible liquids, most
preferably oil and water, are dispersed and stabilized with the addition of a surfac-
tant. A transparent, isotropic, stable, and low-viscous system with water droplets
will be obtained in which the water droplets are surrounded by the surfactant and
stabilized. These water droplets act as nanoreactors for further chemical reactions
and help to control the size of nanoparticles. For instance, ZnO-ZnWO4 nanoparti-
cles were synthesized by emulsion method in which sodium tungstate dehydrate
(Na2WO4·2H2O) and zinc (II) nitrate (Zn (NO3)2·4H2O) were used as metal precur-
sors which were mixed in distilled water to form one microemulsion (sample-A)
[53]. On the other hand, cyclohexane was used as oil agent, and brij35 and 1-butanol
as surfactant and cosurfactant were mixed along with the addition of ammonium
hydroxide (NH4OH) which acts as a precipitating agent to form the second micro-
emulsion (sample-B). A precipitate was obtained with the addition of sample-B to
sample-A in a dropwise manner which was then washed with ethanol and dried in
oven and was finally calcined. The obtained nanoparticles exhibited hexagonal-­
shaped morphology with average diameter of 25 nm. Other examples of microemul-
sion method of synthesis includes development of TiO2- and Zr-doped TiO2
nanoparticles for methylene blue dedradation [54], spinel-type ferrite nanoparticles
for photocatalytic water splitting [55], and Bi2MoO6 nanoparticles for dye degrada-
tion [56].
Sonochemical method is a facile simple approach for the synthesis of photocata-
lytic nanostructures. Ultrasonic waves are used for the chemical reactions that are
carried out during the synthesis. For example, Cu/ZnO/Al2O3 ternary nano-hybrids
were prepared via sonochemical method [57]. Firstly, 1 mmol of Zn(NO3)2·4H2O,
Al(NO3)3·9H2O, and N2H4·H2O were dissolved in water and stirred for a certain
period. The resulting solution was subjected to the ultrasonication for 10 min fol-
lowed by the addition of Cu(NO3)2·6H2O solution in a dropwise manner. The ultra-
sonication was carried for another 5 min, and finally the resultant precipitant was
annealed under vacuum, and the observed morphology was polyhedron microstruc-
tures with diameter size between 100 and 400 nm. ZnO-rGO nanocomposites [58],
FeVO4 nanostructures [59], Tl4CdI6 nanostructures [60], and MnWO4/TmVO4
hybrid nanostructures [61] were also prepared with sonochemical method for vari-
ous photocatalytic activities.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 11

1.4  Applications

1.4.1  Dye Degradation

Various kinds of dyes are being released into water by different industries such as
paper, textile, food processing, cosmetics, and paints. These dyes contaminate the
water and are proven to be hazardous to environment and health. Photocatalysis has
been employed for dye degradation applications from the past few decades, and
various dyes such as acid dyes, basic dyes, direct dyes, pigment dyes, reactive dyes,
etc. have been degraded by using different photocatalytic nanostructures. One such
example was the use of CdS/ZnO heterojunction with nonuniform spherical-like
structure, for the photodegradation of methylene blue in an aqueous solution under
solar irradiation [62].
The degradation mechanism is shown in Fig.  1.5(c). Upon light irradiation,
electron-­hole pairs are generated, and the excited electrons from the conduction
band (CB) of CdS migrate to the CB of ZnO, but the transfer of holes does not take
place as the valence band (VB) of CdS is more cathodic than ZnO. The electrons of

Fig. 1.5 (a) Photocatalytic degradation of methylene blue using pure ZnO and CdS/ZnO. (b)
Recycling photocatalytic tests of CdS/ZnO. (c) Schematic representation of CdS/ZnO photocata-
lytic dye degradation
12 Y. V. Divyasri et al.

ZnO produce superoxide radicals which react with holes and water producing
large  number of hydroxyl radicals which are responsible for the degradation of
methylene blue. The degradation efficiencies of pure ZnO and CdS/ZnO are shown
in Fig. 1.5(a, b). After an irradiation time of 240 min, pure ZnO thin film showed
an efficiency of 54%, whereas the CdS/ZnO heterojunction showed 91% degrada-
tion efficiency. The degradation efficiencies of various photocatalysts are provided
in Table. 1.1.

1.4.2  Pharmaceutical Pollutant Degradation

Rise in the pharmaceutical pollutants in the water is another major concern for the
environment, and photocatalysis is believed to be one of the major tools for degra-
dation of these dangerous pollutants. Magnetic ZnO@g-C3N4 heterojunction is one
among the many photocatalysts that were explored for pharmaceutical pollutant
degradation [74]. ZnO@g-C3N4 is employed for the degradation of sulfamethoxa-
zole (SMX) antibiotic under UV irradiation (Fig. 1.6(a)). The proposed mechanism
is shown in Fig. 1.6(b).
The CB edge potentials of g-C3N4 and ZnO was found to be −1.12  eV
and −0.31 eV, respectively, from which it is clear that the CB of g-C3N4 is more
negative than the ZnO which leads to the migration of excited electrons from CB of
g-C3N4 to ZnO and from there to Fe3O4 upon the UV light irradiation. From there,
the electrons participate directly in pollutant degradation. As it is clear from
Fig. 1.6(b) that the oxidation potential of holes in the VB of g-C3N4 is quite low to
produce hydroxyl radicals, the redox potential of ZnO helps for the production of

Table 1.1  Photocatalytic degradation of various dyes


Degradation time Efficiency of
Photocatalyst Dye (min) degradation (%) Ref.
MoS2-FeZnO MB 140 95.2 [63]
WO3/g-C3N4 MB 90 95 [64]
g-C3N4/MnV2O6 MB/indigo carmine 210/60 95/94 [65]
TiO2-CdO-Ag MB 75 92 [24]
Cd-doped CeO2 Rhodamine B/CR 360 86.42/92.53 [22]
Nd2Zr2O7 Erythrosine/ 50 88/84 [66]
Eriochrome/black T
Fe2V4O13/ZnO Methyl orange 60 99 [67]
TiO2 Reactive black 5 180 70 [68]
CeO2 Reactive violet 1 180 99.9 [69]
SrTiO3/BiOI Crystal violet 30 92.5 [70]
Ag3PO4@MWCNTs@ Malachite green 10 100 [71]
Cr:SrTiO3
Fe@MoPO Malachite green 60 77 [72]
TiO2/rGO Acid Red 14 120 96.38 [73]
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 13

Fig. 1.6 (a) Degradation efficiency of sulfamethoxazole with various photocatalysts. (b) Pollutant
degradation schematic representation

o2 26nm
(a) MTP
(b) 1
e- e- e- e- o2- • 53nm
Concentration (C/Co)

Degradation products 0.8 80nm

0.6 106nm
Excitation Recombination
0.4

H2O or OH– 0.2


h+ h+ h+ h+
MTP
• OH
MTP Degradation products 0
0 30 60 90 120
Degradation products Irradiation Time (min)

Fig. 1.7 (a) Schematic representation of degradation of metoprolol by photocatalysis. (b)


Degradation efficiency by different sizes of TiO2 nanotubes

hydroxyl radicals, and holes can also participate directly in the degradation of pol-
lutant or the intermediate products formed. Accordingly, 90.4% SMX pollutant deg-
radation efficiency was found after 60-min irradiation of UV light which met the
predicted value of 92.51% obtained by response surface methodology (RSM).
Another example of photocatalysis application in pollutant degradation is TiO2
nanotube arrays which are used for the degradation of metoprolol under UV light
irradiation [75]. Degradation mechanism is schematically represented in Fig. 1.7(a).
The degradation of metoprolol was carried out by four different diameter sizes of
TiO2 nanotubes. The degradation efficiency curves are represented in Fig. 1.7(b).
Degradation efficiency was increased from 63.66% to 82.88% with increase in the
nanotubes size from 26 to 53 nm, but further increase of nanotubes to 80–106 nm
does not show any particular increase in the degradation efficiency. The reason for
this is that as the diameter size of the nanotubes increases, the surface area decreases
which affects the absorption ability of TiO2. Further different pharmaceutical pol-
lutant degradation by various photocatalysts is showed in Table. 1.2.
14 Y. V. Divyasri et al.

Table 1.2  various pollutant degradation efficiencies of different photocatalysts with respective
degradation time
Degradation time Efficiency of
Photocatalyst Pollutant (min) degradation (%) Reference
TiO2 Ibuprofen 240 100 [76]
Mg-doped Caffeine 70 98.9 [77]
ZnO-Al2O3
Pt-TiO2-Nb2O5 Diclofenac 20 100 [78]
Ketoprofen 30
Bi2WO6/Fe3O4/GSC Ampicillin 60 95 [79]
Oxytetracycline 94
WO3 Amoxicillin 180 99.9 [80]
N-TiO2/rGO Tetracycline 60 98 [81]
hydrochloride
β-Bi2O3@g-C3N4 Tetracycline 50 80.2 [82]
WO3–TiO2 @ Aspirin 90 98 [83]
gC3N4 Caffeine 97
g-C3N4/Ag/AgCl/ Ibuprofen 60 94.7 [84]
BiVO4
Ag@BiPO4/BiOBr/ Norfloxacin 90 98.1 [85]
BiFeO3
CuBi2O4/Bi2WO6 Tetracycline 60 93 [86]
GO–Ag–ZnFe2O4 17 α-Ethinyl 240 80 [87]
estradiol
RGO-Ce/WO3 Ciprofloxacin 60 100 [88]

1.4.3  Plastic Degradation

Plastic has been used widely in day-to-day life because of its various advantages
such as durability, light weight, and cost-effectiveness. With enormous increase in
the usage caused an enormous increase in disposal of this material which is becom-
ing an environmental threat as the plastic is not biodegradable and also possesses
hydrophobic nature which leads to the formation of fungi and microorganism on its
surface. Even though strategies for plastic degradation such as recycling and bury-
ing are suggested, they also possess certain drawbacks that limited their implemen-
tations. Hence, the need for degradable plastic is a hot topic among the research
fraternity which led the way to the exploration of photocatalysts for the develop-
ment of degradable plastic. The incorporation of photocatalysts in the plastic mate-
rials will help for the degradation when exposed to irradiation as the radicals
produced during the photocatalytic process will react with the organic materials
and lead to oxidation and decomposition. In this strategy of using photocatalysts
for  degradable plastic, TiO2 has been explored extensively due to its various
properties  such as high stability, nontoxicity, high photocatalytic ability, and
­
inexpensiveness.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 15

In this direction, TiO2-based nanocomposites were used for photodegradation


study of polypropylene (PP) under solar irradiation. Reduced graphene oxide
(rGO)–TiO2 composites were used to form uniform PP/TiO2–rGO film whose syn-
thesis is shown in Fig. 1.8(a, b) [89]. The mechanism of photocatalytic degradation
is schematically represented in Fig. 1.8(c). As rGO is a good electron acceptor, the
photogenerated electrons are transferred from the CB of TiO2 to CB of rGO which
helps for the suppression of undesired recombination of electron-hole pairs. Also
the surface area of the film has been increased with the incorporation of rGO which
provides more active sites which in turn improves the photocatalytic activity. The
FESEM characterization helped toward determining of the photocatalyst before and
after photodegradation process, which is presented in Fig. 1.9.
The upper section of the image represents the morphology of PP/TiO2 film, while
the downward section is of PP/TiO2–rGO. Figure 1.9(a1) and (b1) is before irradi-
ance, which shows a smooth surface, while in Fig. 1.9(a3), (a4), (b3), and (b4), it
can be seen that there are deep cavities formed on the surface which show that the
irradiation has a significant influence on the films and thus the degradation of poly-
propylene has been confirmed. It has been determined that the degradation effi-
ciency of PP/TiO2–rGO is better than PP/TiO2.
Another example is the use of ZnO for the degradation of low-density polyethyl-
ene (LDPE) under UV irradiation [90]. Firstly, ZnO nanoparticles were grafted with
polystyrene to prepare ZnO composites, and then LDPE film with ZnO nanocom-
posites were prepared. The LDPE-ZnO film was exposed to UV irradiation for
200 h. The weight loss assessment and tensile strength of pure LDPE and grafted
ZnO-LDPE for various irradiation times of 50, 100, and 200  min with different
concentration of ZnO were investigated, which helped for the confirmation of pho-
todegradation. With the increasing ZnO concentration, the weight loss also
increased, but after a certain amount, there is a decrease which states that the right
amount of ZnO incorporation should be done carefully.

Fig. 1.8 (a) Synthesis of TiO2–rGO nanocomposite. (b) Synthesis of uniform PP/TiO2–rGO film.
(c) Schematic representation of photocatalytic degradation
16 Y. V. Divyasri et al.

Fig. 1.9  FESEM images of PP/TiO2 and PP/TiO2–rGO before and after irradiance of light

Fig. 1.10  The appearance of film samples before and after underwent photo- and biodegradation
studies [91]

Other examples include TiO2 nanoparticles in combination with biodegradable


polymer polylactic acid for degradation of LDPE as shown in Fig. 1.10 [91], photo-
degradation of polypropylene-­ascorbic acid- TiO2 composite films [92], photodeg-
radation of polyvinyl borate (PVB)-TiO2 nanocomposites prepared by condensation
of polyvinyl alcohol and boric acid in the presence of TiO2 nanoparticles [93], and
ZnO nanorods used for degradation of LDPE under visible light [94].
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 17

1.4.4  CO2 Reduction

Rapid consumption of fossil fuels leads to the increment of CO2 in the atmosphere
which is one of the main contributors to the greenhouse effect. Photocatalytic CO2
conversion is a greener approach to convert CO2 into carbon-containing fuels to
tackle this environmental issue. In this process, various photocatalysts have been
exploited for conversion of CO2 into CO, HCOOH, CH3OH, and CH4. Table  1.3
represents photocatalytic conversion of CO2 into various fuels using various photo-
catalytic systems.
A direct Z-scheme TiO2/CuInS2 nanostructure was used for the reduction of CO2
into CH4 and CH3OH [95]. Firstly, TiO2 nanofibers were synthesized by electrospun
method followed by the development of TiO2/CuInS2 heterostructure through
hydrothermal process. The CO2 reduction efficiency and mechanism are shown in
Fig. 1.11(a, b). With light irradiation, electron-hole pairs are generated in the CB
and VB of both TiO2 and CuInS2. TiO2 possess more negative CB edge potential
than that of CuInS2. This results in the recombination of electrons from the CB of
TiO2 with the holes in the VB of CuInS2. Electrons in the CB of CuInS2 with high
reducibility are used for the reduction of CO2–CH4 and CH3OH, while holes in the
VB of TiO2 with high oxidizibility react with water molecules. Type-II heterojunction-­
based ZnO/ZnSe composite is also used for the reduction of CO2 to methanol whose
reduction mechanism and rate of reduction are shown in Fig. 1.12(a, b) [96]. The
electrons from the CB of ZnSe transfer to CB of ZnO which reduced the CO2 mol-
ecules absorbed on the surface of photocatalyst, while the holes in the VB of ZnSe
are used for oxidation of C3H8O to form C3H6O.

Table 1.3  Photocatalytic CO2 reduction by various photocatalysts


Photocatalyst Final product Yield in μmol h−1 g−1 Reference
g-C3N4/FeWO4 CO 6 [97]
α-Fe2O3/g-C3N4 CO 27.2 [98]
ZnO/Ag1−xCux/CdS CO 327.4 [99]
BVO/C/Cu2O CO 3.01 [100]
Ag/Pd/TiO2 CH4 79 [101]
TiO2/Ti3C2 CH4 0.22 [102]
In2O3 coated with carbon CO/CH4 126.6/27.9 [103]
In2O3/CeO2/HATP CO/CH4 32.03/16.94 [104]
MoS2/TiO2 CO/CH4 269.97/49.93 [105]
MXene/Bi2WO6 CH4/CH3OH 1.78/0.44 [106]
TiO2–MnOx–Pt CH4/CH3OH 104/91 [107]
18 Y. V. Divyasri et al.

Build-in
Reduction of CO2 (µmol h-1 g-1)
2.5 (a) (b) electric
CH4 -1 TiO2 field
CH3OH e- e- e- e- CO2

V vs NHE pH = 0 (eV)
2.0
CH4 and
0 - +
1.5 CH3OH
e-
1 - +
1.0 h+
- +
0.5 2 CulnS2
- +
0.0 3 H O h+ h +
T TC1 TC2.5 TC5 TC10 2
• OH
Samples

Fig. 1.11 (a) CO2 reduction of TiO2/CuInS2 with various concentrations. (b) Z-scheme photocata-
lytic reduction mechanism

Fig. 1.12 (a) Photocatalytic CO2 reduction mechanism of ZnO/ZnSe. (b) CO2 conversion effi-
ciency into methanol by various photocatalysts

1.4.5  N2 Fixation

Nitrogen (N2) fixation through photocatalysis involves the reduction of N2 to ammo-


nia (NH3), which plays a crucial role in the biological synthesis of agricultural fertil-
izers. Even though NH3 is synthesized through traditional Haber-Bosch process,
disadvantages such as harsh reaction conditions and high energy consumption make
it necessary to find a cost-effective and environmentally friendly approach for NH3
synthesis. Photocatalysis promises both these advantages and hence has been
explored widely in the recent times. The reduction reaction for N2 fixation is similar
to that of CO2 reduction, but the main limitation in the process of photocatalytic N2
fixation is the adsorption of N2 molecules onto the surface of the photocatalyst.
Bismuth subcarbonate (BOC) with controllable defect density (BOC-X) was
synthesized using sodium bismuthate (NaBiO3) and graphitic carbon nitride
(g-C3N4) as precursors via hydrothermal process in order to overcome this draw-
back [108]. The photocatalytic N2 fixation is carried under three different irradiation
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 19

Fig. 1.13 N2 fixation efficiencies of various BOC-X samples under (a) UV irradiation, (b) visible
light irradiation, and (c) sunlight irradiation. (d) Schematic representation of photocatalytic N2
fixation mechanism under two different light sources

sources such as UV, visible light, and solar irradiation. The N2 conversion efficien-
cies with various concentrations of g-C3N4 deposited onto the defect-engineered
BOC-X are shown in Fig. 1.13(a–c). It can be seen from Fig. 1.13(a) the samples
showed highest NH3 yield under UV light irradiation. The photocatalytic mecha-
nism under two different light sources is shown in Fig. 1.13(d).
The defects created on the surface of photocatalysts provide oxygen vacancies
which play a major role in the absorption of N2 molecules onto the surface, thereby
increasing the photocatalytic efficiency. The excited electrons transfer from VB of
BOC-X to the CB when the photon energy is higher than that of the bandgap of
photocatalyst and participate in the conversion of N2 to NH3 directly. But when the
photon energy is less than the bandgap, the excited electrons migrate to the defect
levels where they participate in the N2 reduction reactions. There was a decrease in
the NH3 yield after BOC-2 which states that the proper defect engineering is highly
important. Too many of defects in the system will require higher photon energy for
excitation of electrons, while less number of defects will result in the decrease in
reduction ability of electrons, both of which are not favorable for better N2 fixation
(Table 1.4).

1.4.6  Heavy Metal Reduction

One of the main causes of water contamination is the continuous ingestion of heavy
toxic metals like chromium (Cr), copper (Cu), manganese (Mn), nickel (Ni), lead
(Pb), silver (Ag), and cadmium (Cd) into the water in the form of industrial wastage.
20 Y. V. Divyasri et al.

Table 1.4 NH3 yield obtained by photocatalytic N2 fixation using various photocatalysts


Yield of NH3 Light Irradiation time
Photocatalyst (μmol g−1 h−1) source (min) Ref.
ZnIn2S4/BiOCl 14.6 Visible 360 [109]
light
A-CeOx 109 Visible 60 [110]
light
TiO2@C/g-C3N4 250.6 Visible 120 [111]
light
KOH-treated g-C3N4 3632 Visible 240 [112]
light
BiO quantum dots 1226 Sunlight 24 h [113]
TiO2 nanotubes 106.6 Sunlight 60 [114]
TiO2 – x 2.5 UV 48 h [115]
g-C3N4 codoped with 6.2 mg L−1 h−1 g−1 Visible 240 [116]
sulfur light
g-C3N4/Ag2CO3 11 mg L−1 h−1 g−1 Visible 240 [117]
light
Au–Ag2O 28.2 mg m−2 h−1 Sunlight 60 [118]
Ru/TiO2(P25) 419 μmol L−1 g−1 UV–Vis 480 [119]
CdS nanorods-NiS 1.7 mg L−1 Visible 60 [120]
light

Reduction of these heavy toxic metals into nontoxic or less toxic elements through
photocatalysis is considered as a greener and cost-effective approach. Various pho-
tocatalysts are employed for the reduction of hexavalent chromium Cr (VI), one of
the heavy toxic metals, to less toxic Cr (III) through photocatalysis. One such inves-
tigation involved the usage of CC@SnS2/SnO2 heterojunction for the reduction of
Cr (VI) under visible light irradiation as shown in Fig. 1.14(a–d) [121].
Upon light irradiation, electrons and holes are generated in the CB and VB of
SnS2, but no generation of charge carriers occurs in SnO2 as it is incapable of absorb-
ing visible light. Excited electrons migrate to the CB of SnO2, while the holes
remain in the VB of SnS2 leading to the suppression of recombination of charge
carriers. Finally the reduction of Cr (VI) to Cr (III) was carried out by the reaction
of electrons in the CB of SnO2 with the Cr (VI) absorbed on the surface, and the
holes in the VB of SnS2 participate in the oxidation of H2O producing O2. Reduction
activity was carried out for different CC@SnS2/SnO2 composites prepared at differ-
ent calcination time. Highest photocatalytic activity was shown by C ­ C@SnS2/
SnO2-2 at pH-2, which clearly shows that the photocatalyst size and pH value play
an important role in the reduction performance. As the size of the nanosheets
increased, the absorption ability of the photocatalyst decreased leading to the poor
reduction performance. Some of the photocatalysts developed for the heavy metal
reduction are listed in Table 1.5.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 21

Fig. 1.14 (a) Cr (VI) removal efficiency of CC@SnS2/SnO2-2 at different pH values. (b) Removal
efficiencies of different CC@SnS2/SnO2-x composites. (c) Cyclic performance of CC@SnS2/
SnO2-2. (d) schematic representation of photocatalytic reduction mechanism

Table 1.5  Heavy metal reduction efficiencies using various photocatalysts


Reduction
Photocatalyst Heavy metal efficiency% Light source Reference
ZnO Cu (II), Ag (I), Pb (II) >85 Visible and [122]
Cr (VI), Mn (II), Cd (II), <15 UV
Ni (II)
Fe3O4/RGO/ Ce (III) 66.3 Visible light [123]
PAM Cu (II) 34.8
Ag (I) 52.4
Cd (II) 52.9
TiO2/graphene Cd (II) 66.32 UV [124]
Pb (II) 88.96
TiO2 nanofibers Cr (VI) 90 Solar [125]
irradiation
Nb2O5-CF Cr (VI) 99.9 UV–Vis [126]
Sulfur-doped Cr (VI) 100 Visible light [127]
g-C3N4
BiOBr/Ti3C2 Cr (VI) 100 UV–Vis [128]
22 Y. V. Divyasri et al.

1.4.7  Antimicrobial Applications

Disinfection of contaminated water by killing dangerous microorganisms such as


bacteria and other pathogens is one of the important applications of photocatalysis.
Inactivation of pathogens such as Escherichia coli (E. coli), Streptococcus mutans,
Staphylococcus aureus, and Porphyromonas gingivalis has been carried out by
employing various photocatalysts. The highly reactive species that are activated
during the photocatalysis damage the cell walls of microorganisms first and then the
inner layers of the cell, thereby causing the leakage of potassium ions that affects
the functionality of the cell finally leading to the death of microorganism.
Ag/PCN nanocomposites were investigated for the photocatalytic antimicrobial
applications against E. coli, where silver (Ag) nanoparticles were decorated with
phosphorus-doped g-C3N4 (PCN) under visible light irradiation [129]. The reduc-
tion mechanism shown in Fig. 1.15(a) involves the migration of excited electrons to
Ag (good electron accepter) where they interact with the oxygen and generate O2−
which react with the bacteria absorbed on the surface thus by killing the E. coli. On
the other hand, holes in the VB of PCN cannot participate in the oxidation of H2O
due to higher negative VB potential (1.57  eV) than the required redox potential
(1.99 eV); hence, these holes directly react and help in the inactivation of E. coli
bacteria. From the graph shown in Fig. 1.15(b), the inactivation of E. coli by g-C3N4
and PCN was extremely low with just 1.2 log and 1.5 log decrease in density of cell
after an irradiation time of 60 min, whereas Ag/PCN showed excellent photocata-
lytic ability with complete inactivation of cells. E. coli bacteria inactivation is the
most widely explored antimicrobial application by using distinct photocatalytic sys-
tems which are presented in Table 1.6. Examples of other bacterial and fungal dis-
infection tests carried out by well-diffusion method and disk-diffusion methods are
presented in Tables 1.7 and 1.8.

Fig. 1.15 (a) Photocatalytic mechanism for disinfection of E. coli by Ag/PCN. (b) Time-kill activ-
ity of Ag/PCN under visible light irradiation
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 23

Table 1.6 Microbial Photocatalyst Irradiation time (min) Reference


inactivation of E. coli bacteria
Ag/CeO2-M 150 [130]
under visible light irradiation
using various photocatalysts Bi2MoO6-AgBr 90 [131]
VS4/CP 30 [132]
Ag/AgBr/ZnFe2O4 120 [133]
Ag/ZnO/g-C3N4 120 [134]
AgBr/g-C3N4 150 [135]
Cu-TiO2 240 [136]
CdSe QDs/graphene/TiO2 60 [137]
Se-N-codoped TiO2 60 [138]

Table 1.7  Antimicrobial inactivation by well-diffusion method using CdS–Ag2S photocatalyst [139]
Reduction rate (%)
Photocatalyst P. aeruginosa E. coli S. aureus
CdS 66 ± 3.2 80.5 ± 3.2 45 ± 1.5
Ag2S 55.1 ± 1.5 58.7 ± 2.7 41.5 ± 1.9
5% CdS–Ag2S 70.3 ± 2.3 82.3 ± 3.4 46.5 ± 2.7
10% CdS–Ag2S 80.3 ± 2.4 90.7 ± 2.3 53.8 ± 2.3
15% CdS–Ag2S 74.1 ± 4.2 86.4 ± 1.8 49.6 ± 1.5

Table 1.8 Antimicrobial inactivation by disk-diffusion method using MnO2/CdTiO3


photocatalyst [140]
Zone of inhibition (mm)
Antimicrobial model Standard antibiotic disk MnCdTi-0 MnCdTi-1 MnCdTi-2
Fungi
A. flavus 10.11 ± 0.44 3.12 ± 0.14 10.21 ± 0.10 15.11 ± 0.13
C. albicans 9.54 ± 0.10 2.34 ± 0.31 7.21 ± 0.31 10.14 ± 0.10
Bacteria
E. coli 18.31 ± 0.23 8.12 ± 0.11 12.34 ± 0.54 15.11 ± 0.31
S. aureus 22.30 ± 0.22 10.05 ± 0.23 15.31 ± 0.45 18.32 ± 0.11

1.5  Conclusions and Outlook

Photocatalysis is a cost-effective and environmentally friendly technique that uti-


lizes the solar radiation as energy source and used as an effective way of solving
energy crisis and several environmental issues. Photocatalysis efficiency highly
depends on the type of photocatalyst that is used and various “nanostructured”
semiconductors such as TiO2, g-C3N4, ZnO, CeO2, WO3, etc., thanks to their advan-
tages such as high surface area and tunable physiochemical properties. Eventually,
24 Y. V. Divyasri et al.

these photocatalytic nanostructures show excellent efficiency especially in the envi-


ronmental applications such as CO2 reduction, N2 fixation, heavy toxic metal reduc-
tion, antimicrobial inactivation, and also degradation of plastic, dyes, and
pharmaceutical pollutants.
Even though these photocatalytic nanostructures showed promising efficiency in
photocatalytic activities due to certain limitations and drawbacks, its utilization to a
full extent and on a large scale is being axed. One of the major factors that contrib-
ute in the reduction of photocatalytic activity is the bandgap of photocatalyst nano-
structures used. Most of the photocatalysts explored contain higher bandgaps
limiting the photocatalytic activity to be carried out in the UV region thus by limit-
ing the utilization of solar light. Hence more focus should be exerted on the devel-
opment of nanostructures with appropriate bandgap to make use of renewable solar
energy. Second factor is the fast recombination of charge carriers. Various strategies
such as doping, formation of heterojunction, and Z-scheme have been proposed to
overcome these problems. Doping of the cocatalyst should be done carefully as the
doping beyond optimum level will lead to the reduction of active sites thus by
reducing the light harvesting ability of photocatalyst. Adsorption of the target ele-
ment onto the surface of the photocatalyst is one more limitation that needs to be
focused on. One of the solutions proposed to this problem was the defect engineer-
ing particularly generating oxygen vacancies. But more investigation is needed as
too much increase or decrease in number of vacancies in the nanostructure will
highly affect the photocatalytic activity. Size of the nanostructures should be main-
tained carefully as increase in the size will affect the surface area of the photocata-
lyst which will reduce the number of active sites leading to reduction of photocatalytic
efficiency. In conclusion, more investigation and research has to be done for the
development of cost-effective and visible light-responsive “nanostructures,” which
are very much needed as the photocatalytic nanostructures are proven to be highly
appreciable for the environmental applications.

Acknowledgments  One of the authors, MS, gratefully acknowledges the Department of Science
and Technology, Govt. of India, for the funding support through the DST-INSPIRE Faculty Award
[DST/INSPIRE/04/2016/002227, 14-02-2017].

References

1. Chae, Y., & An, Y.  J. (2018). Current research trends on plastic pollution and ecological
impacts on the soil ecosystem: A review. Environmental Pollution, 240, 387–395.
2. Patel, M., Kumar, R., Kishor, K., Mlsna, T., Pittman, C.  U., & Mohan, D. (2019).
Pharmaceuticals of emerging concern in aquatic systems: Chemistry, occurrence, effects, and
removal methods. Chemical Reviews, 119, 3510–3673.
3. Kumar, V., Sharma, A., Kaur, P., Sidhu, G. P. S., Bali, A. S., Bhardwaj, R., Thukral, A. K., &
Cerda, A. (2019). Pollution assessment of heavy metals in soils of India and ecological risk
assessment: A state-of-the-art. Chemosphere, 216, 449–462.
4. Byrnea, C., Subramanian, G., & Pillai, S. C. (2018). Recent advances in photocatalysis for
environmental applications. Journal of Environmental Chemical Engineering, 6, 3531–3555.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 25

5. Ahmed, S. N., & Haider, W. (2018). Heterogeneous photocatalysis and its potential applica-
tions in water and wastewater treatment: A review. Nanotechnology, 29, 342001.
6. Fujishima, A., & Honda, K. (1972). Electrochemical photolysis of water at a semiconductor
electrode. Nature, 238, 37–38.
7. Jardim, W. F., Moraes, S. G., & Takiyama, M. M. K. (1997). Photocatalytic degradation of
aromatic chlorinated compounds using TiO2: Toxicity of intermediates. Water Research, 31,
1728–1732.
8. Martin, S.  T., Lee, A.  T., & Hoffmann, M.  R. (1995). Chemical mechanism of inorganic
oxidants in the TiO2/UV process: Increased rates of degradation of chlorinated hydrocarbons.
Environmental Science & Technology, 29, 2567–2573.
9. Maness, P. C., Smolinski, S., Blake, D. M., Huang, Z., Wolfrum, E. J., & Jacoby, W. A. (1999).
Bactericidal activity of photocatalytic TiO2 reaction: Toward an understanding of its killing
mechanism. Applied and Environmental Microbiology, 65, 4094–4098.
10. Miklos, D.  B., Remy, C., Jekel, M., Linden, K.  G., Drewes, J.  E., & Hübner, U. (2018).
Evaluation of advanced oxidation processes for water and wastewater treatment: A critical
review. Water Research, 139, 118–131.
11. Poza-Nogueiras, V., Rosales, E., Pazos, M., & Sanroman, M. A. (2018). Current advances
and trends in electro-Fenton process using heterogeneous catalysts – a review. Chemosphere,
201, 399–416.
12. Du, J., Wang, Z., Li, Y. H., Li, R. Q., Li, X. Y., & Wang, K. Y. (2019). Establishing WO3/g-­
C3N4 composite for “Memory” photocatalytic activity and enhancement in photocatalytic
degradation. Catalysis Letters, 149, 1167–1173.
13. Low, J., Yu, J., Jaroniec, M., Wageh, S., & Al-Ghamdi, A. A. (2017). Heterojunction photo-
catalysts. Advanced Materials, 24, 1601694.
14. Xu, Q., Zhang, L., Yu, J., Wageh, S., Al-Ghamdi, A.  A., & Jaroniec, M. (2018). Direct
Z-scheme photocatalysis: Principles, synthesis and applications. Materials Today, 22,
1042–1063.
15. Wang, D., Pillai, S. C., Ho, S. H., Zeng, J., Li, Y., & Dionysiou, D. D. (2018). Plasmonic-­
based nanomaterials for environmental remediation. Applied Catalysis B: Environmental,
237, 721–741.
16. Reddy, N. L., Rao, V. N., Vijayakumar, M., Santhosh, R., Anandan, S., Karthik, M., Shankar,
M. V., Reddy, K. R., Shetti, N. P., Nadagouda, M. N., & Aminabhavi, T. M. (2019). A review
on frontiers in plasmonic nano-photocatalysts for hydrogen production. International Journal
of Hydrogen Energy, 44, 10453–10472.
17. N.R. Yogamalar, S. Kalpana, V. Senthil, A. Chithambararaj, Ferroelectrics for photocatalysis,
multifunctional photocatalytic materials for energy, 2018, pp. 3017–324.
18. Han, Q., Wang, B., Gao, J., Cheng, Z., Zhao, Y., Zhang, Z., & Qu, L. (2016). Atomically thin
mesoporous nanomesh of graphitic C3N4 for high-efficiency photocatalytic hydrogen evolu-
tion. ACS Nano, 10, 2745–2751.
19. Jiang, L., Yuana, X., Pana, Y., Liang, J., Zeng, G., Wu, Z., & Wang, H. (2017). Doping of
graphitic carbon nitride for photocatalysis: A review. Applied Catalysis B: Environmental,
217, 388–406.
20. Nasirian, M., Lin, Y.  P., Bustillo-Lecompte, C.  F., & Mehrvar, M. (2018). Enhancement
of photocatalytic activity of titanium dioxide using non-metal doping methods under vis-
ible light: A review. International Journal of Environmental Science and Technology, 15,
2009–2032.
21. Deebasree, J.  P., Maheskumar, V., & Vidhya, B. (2018). Investigation of the visible light
photocatalytic activity of BiVO4 prepared by sol-gel method assisted by ultrasonication.
Ultrasonics Sonochemistry, 45, 123–132.
22. Gnanam, S., & Rajendran, V. (2018). Facile sol–gel preparation of Cd-doped cerium oxide
(CeO2) nanoparticles and their photocatalytic activities. Journal of Alloys and Compounds,
735, 1854–1862.
26 Y. V. Divyasri et al.

23. Alam, U., Shah, T. A., Khan, A., & Muneer, M. (2019). One-pot ultrasonic assisted sol-gel
synthesis of spindle-like Nd and V codoped ZnO for efficient photocatalytic degradation of
organic pollutants. Separation and Purification Technology, 212, 427–437.
24. Azimi-Fouladia, A., Hassanzadeh-Tabrizia, S. A., & Saffar-Teluri, A. (2018). Sol–gel syn-
thesis and characterization of TiO2–CdO–Ag nanocomposite with superior photocatalytic
efficiency. Ceramics International, 44, 4292–4297.
25. Mazloom, F., Ghiyasiyan-Aran, M., Monsef, R., & Salavati-Niasari, M. (2018). Photocatalytic
degradation of diverse organic dyes by sol–gel synthesized Cd2V2O7 nanostructures. Journal
of Materials Science: Materials in Electronics, 29, 18120–18127.
26. Yu, H., Wang, J., Xia, C., Yan, X., & Cheng, P. (2018). Template-free hydrothermal synthesis
of Flower-like hierarchical zinc oxide nanostructures. Optik, 168, 778–783.
27. Li, H., Meng, F., Gong, J., Fan, Z., & Qin, R. (2018). Template-free hydrothermal synthe-
sis, mechanism and photocatalytic properties of Core–Shell CeO2 nanospheres. Electronic
Materials Letters, 14, 474–487.
28. Xu, M., Jia, S., Cheng, C., Zhang, Z., Yan, J., Guo, Y., Zhang, Y., Zhao, W., Yun, J., & Wang,
Y. (2018). Microwave-assistant hydrothermal synthesis of SnO2@ZnO hierarchical nano-
structures enhanced photocatalytic performance under visible light irradiation. Materials
Research Bulletin, 106, 74–80.
29. Prabhu, Y. T., Rao, V. N., Shankar, M. V., Sreedhar, B., & Pal, U. (2019). Facile hydrothermal
synthesis of CuO@ZnO heterojunction nanostructures for enhanced photocatalytic hydrogen
evolution. New Journal of Chemistry, 43, 6794–6805.
30. Zhou, B., Li, Y., Bai, J., Li, X., Li, F., & Liu, L. (2019). Controlled synthesis of rh-In2O3
nanostructures with different morphologies for efficient photocatalytic degradation of oxy-
tetracycline. Applied Surface Science, 464, 115–124.
31. Jaihindh, D. P., Thirumalraj, B., Chen, S. M., Balasubramanian, P., & Fu, Y. P. (2019). Facile
synthesis of hierarchically nanostructured bismuth vanadate: An efficient photocatalyst for
degradation and detection of hexavalent chromium. Journal of Hazardous Materials, 367,
647–657.
32. Liu, Y., Ma, Y., Liu, W., Shang, Y., Zhu, A., Tan, P., Xiong, X., & Pan, J. (2018). Facet and
morphology dependent photocatalytic hydrogen evolution with CdS nanoflowers using a
novel mixed solvothermal strategy. Dalton Transactions, 47, 1325–1336.
33. Mao, Y., Li, Y., Zou, Y., Shen, X., Zhu, L., & Liao, G. (2019). Solvothermal synthesis
and photocatalytic properties of ZnO micro/nanostructures. Ceramics International, 45,
1724–1729.
34. Rao, M. P., Wu, J. J., Asiri, A. M., Anandan, S., & Ashokkumar, M. (2018). Photocatalytic
properties of hierarchical CuO nanosheets synthesized by a solution phase method. Journal
of Environmental Sciences, 69, 115–124.
35. Ebadi, M., Amiri, O., & Sabet, M. (2018). Synthesis of CeO2/Au/Ho nanostructures as
novel and highly efficient visible light driven photocatalyst. Separation and Purification
Technology, 190, 117–122.
36. Zinatloo-Ajabshir, S., Mortazavi-Derazkola, S., & Salavati-Niasari, M. (2017). Preparation,
characterization and photocatalytic degradation of methyl violet pollutant of holmium oxide
nanostructures prepared through a facile precipitation method. Journal of Molecular Liquids,
213, 306–313.
37. Zinatloo-Ajabshir, S., Mortazavi-Derazkola, S., & Salavati-Niasari, M. (2017). Nd2O3 nano-
structures: Simple synthesis, characterization and its photocatalytic degradation of methylene
blue. Journal of Molecular Liquids, 234, 430–436.
38. Liu, W., Yin, L., Zhang, R., Yang, H., Ma, J., & Cao, W. (2018). One-step synthesis of
SnO hierarchical architectures under room temperature and their photocatalytic properties.
Nanotechnology, 29, 28.
39. Zwilling, V., Darque-Ceretti, E., Forveille, A. B., David, D., Perrin, M. Y., & Aucouturier,
M. (1999). Structure and physicochemistry of anodic oxide films on titanium and TA6V
alloy. Surface and Interface Analysis, 27, 629–637.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 27

40. Moriai, K., Nakajima, N., Moriyoshi, C., & Maruyama, H. (2018). Synthesis of TiO2 nano-
tubes: Effect of post-treatment on crystallinity and photocatalytic activity. Materials Research
Express, 4.
41. Brammer, K.  S., Frandsen, C.  J., & Jin, S. (2012). TiO2 nanotubes for bone regeneration.
Trends in Biotechnology, 30, 315–322.
42. Al-Hazeem, N. Z. A. (2018). Nanofibers and electrospinning method, novel nanomaterials—
synthesis and applications. IntechOpen.
43. Wang, K., Zhang, W., Lou, F., Wei, T., Qian, Z., & Guo, W. (2018). Preparation of electrospun
heterostructured hollow SnO2/CuO nanofibers and their enhanced visible light photocatalytic
performance. Journal of Solid State Electrochemistry, 22, 2413–2423.
44. Deng, J., Liu, J., Dai, H., & Wang, W. (2018). Preparation of α-Fe2O3 nanowires through
electrospinning and their Ag3PO4 heterojunction composites with enhanced visible light pho-
tocatalytic activity. Ferroelctrics, 528, 58–65.
45. Shi, R., Zhang, Y., Wang, X., Ma, Q., Zhang, A., & Yang, P. (2018). Electrospun ZnFe2O4
nanotubes and nanobelts: Morphology evolution, formation mechanism and Fenton-like pho-
tocatalytic activities. Materials Chemistry and Physics, 207, 114–122.
46. Pascariu, P., Homocianu, M., Cojocaru, C., Samoila, P., Airinei, A., & Suchea, M. (2019).
Preparation of La doped ZnO ceramic nanostructures by electrospinning–calcination method:
Effect of La3+ doping on optical and photocatalytic properties. Applied Surface Science,
476, 16–27.
47. Campagnolo, L., Lauciello, S., Athanassiou, A., & Fragouli, D. (2019). Au/ZnO Hybrid
nanostructures on electrospun polymeric mats for improved photocatalytic degradation of
organic pollutants. Water, 11, 1787.
48. Mohassel, R., Sobhani, A., Salavati-Niasari, M., & Goudarzi, M. (2018). Pechini synthesis
and characteristics of Gd2CoMnO6 nanostructures and its structural, optical and photocata-
lytic properties. Spectrochim Acta Part A: Mol Biomol Spectrosc, 204, 232–240.
49. Morassaei, M. S., Zinatloo-Ajabshir, S., & Salavati-Niasari, M. (2017). Nd2Sn2O7 nanostruc-
tures: New facile Pechini preparation, characterization, and investigation of their photocata-
lytic degradation of methyl orange dye. Advanced Powder Technology, 28, 697–705.
50. Aflaki, M., & Davar, F. (2016). Synthesis, luminescence and photocatalyst properties of zirco-
nia nanosheets by modified Pechini method. Journal of Molecular Liquids, 221, 1071–1079.
51. Ranjeh, M., Beshkar, F., Amiri, O., Salavati-Niasari, M., & Moayedi, H. (2020). Pechini
sol–gel synthesis of Cu2O/Li3BO3 and CuO/Li3BO3 nanocomposites for visible light-driven
photocatalytic degradation of dye pollutant. Journal of Alloys and Compounds, 815, 152451.
52. Valian, M., Beshkar, F., & Salavati-Niasari, M. (2017). Urchin-like Dy2CoMnO6 double
perovskite nanostructures: New simple fabrication and investigation of their photocatalytic
properties. Journal of Materials Science: Materials in Electronics, 28, 12440–12447.
53. Emsaki, M., Hassanzadeh-Tabrizi, S. A., & Saffar-Teluri, A. (2018). Microemulsion synthe-
sis of ZnO–ZnWO4 nanoparticles for superior photodegradation of organic dyes in water.
Journal of Materials Science: Materials in Electronics, 29, 2384–2391.
54. Mahdi Honarmand, M., Mehr, M. E., Yarahmadi, M., & Siadati, M. H. (2019). Effects of dif-
ferent surfactants on morphology of TiO2 and Zr-doped TiO2 nanoparticles and their applica-
tions in MB dye photocatalytic degradation. SN Applied Sciences, 1, 505.
55. Rodriguez-Rodriguez, A.  A., Moreno-Trejo, M.  B., Melendez-Zaragoza, M.  J., Collins-­
Martinez, V., Lopez-Ortiz, A., Martinez-Guerra, E., & Sanchez-Dominguez, M. (2019).
Spinel-type ferrite nanoparticles: Synthesis by the oil-in-water micro-emulsion reaction
method and photocatalytic water-splitting evaluation. International Journal of Hydrogen
Energy, 44, 12421–124291.
56. Zhang, L., Dai, Z., Zheng, G., Mu, J., & Yao, Z. (2017). Synthesis and photocatalytic
properties of Bi2MoO6 nanoparticles prepared via a water-in-oil microemulsion method.
Ferroelectrics, 530, 17–24.
28 Y. V. Divyasri et al.

57. Mousavi-Kamazani, M. (2019). Facile sonochemical-assisted synthesis of Cu/ZnO/Al2O3


nanocomposites under vacuum: Optical and photocatalytic studies. Ultrasonics: Sonochem,
58, 104636.
58. Yein, W. T., Wang, Q., Feng, X., Li, Y., & Wu, X. (2018). Enhancement of photocatalytic per-
formance in sonochemical synthesized ZnO–rGO nanocomposites owing to effective interfa-
cial interaction. Environmental Chemistry Letters, 16, 251–264.
59. Ghiyasiyan-Arani, M., Salavati-Niasari, M., Masjedi-Arani, M., & Mazloom, F. (2018). An
easy sonochemical route for synthesis, characterization and photocatalytic performance of
nanosized FeVO4 in the presence of aminoacids as green capping agents. Journal of Materials
Science: Materials in Electronics, 29, 474–485.
60. Ghanbari, M., & Salavati-Niasari, M. (2018). Tl4CdI6 nanostructures: Facile sonochemical
synthesis and photocatalytic activity for removal of organic dyes. Inorganic Chemistry, 57,
11443–11455.
61. Sobhani-Nasab, A., Pourmasoud, S., Ahmadi, F., Wysokowski, M., Jesionowski, T., Ehrlich,
H., & Rahimi-Nasrabadi, M. (2019). Synthesis and characterization of MnWO4/TmVO4 ter-
nary nano-hybrids by an ultrasonic method for enhanced photocatalytic activity in the degra-
dation of organic dyes. Materials Letters, 238, 159–162.
62. Velanganni, S., Pravinraj, S., Immanuel, P., & Thiruneelakandan, R. (2018). Nanostructure
CdS/ZnO heterojunction configuration for photocatalytic degradation of Methylene blue.
Physica B: Condensed Matter, 534, 56–62.
63. Ghalajkhani, A., Haghighi, M., & Shabani, M. (2018). Efficient photocatalytic degradation
of methylene blue in aqueous solution over flowerlike nanostructured MoS2-FeZnO stag-
gered heterojunction under simulated solar-light irradiation. Journal of Photochemistry and
Photobiology A: Chemistry, 359, 145–156.
64. Liu, X., Jin, A., Jia, Y., Xia, T., Deng, C., Zhu, M., Chen, C., & Chen, X. (2017). Synergy of
adsorption and visible-light photocatalytic degradation of methylene blue by a bifunctional
Z-scheme heterojunction of WO3/g-C3N4. Applied Surface Science, 405, 359–371.
65. Nithya, M., Vidhya, S., & Keerthi. (2019). Novel g-C3N4/MnV2O6 heterojunction photo-
catalyst for the removal of methylene blue and indigo carmine. Chemical Physics Letters,
731, 136832.
66. Zinatloo-Ajabshir, S., & Salavati-Niasari, M. (2017). Photo-catalytic degradation of eryth-
rosine and eriochrome black T dyes using Nd2Zr2O7 nanostructures prepared by a modified
Pechini approach. Separation and Purification Technology, 179, 77–85.
67. Gowthami, K., Krishnakumar, B., Sobral, A. J. F. N., Thirunarayanan, G., Swaminathan, M.,
Siranjeevi, R., Rajachandrasekar, T., & Muthuvel, I. (2019). Fabrication of hybrid Fe2V4O13/
ZnO heterostructure for effective mineralization of aqueous methyl orange solution. Journal
of Cluster Science.
68. Becerril-Altamirano, N.  L., López, R.  T. H., Reyes, L.  G., Parra, A.  R. S., Lopez, R.  R.,
Jimenez, A. M., & Hernández-Perez, I. Reactive Black-5 photodegradation by TiO2 thin films
prepared by ultrasonic spray. Journal of Physics: Conference Series, 1221, 2019, 012027.
69. Sane, P. K., Tambat, S., Sontakke, S., & Nemade, P. (2018). Visible light removal of reac-
tive dyes using CeO2 synthesized by precipitation. Journal of Environmental Chemical
Engineering, 6, 4476–4489.
70. Xia, Y., He, Z., Su, J., Liu, Y., Tang, B., & Li, X. (2018). Fabrication of novel n-SrTiO3/p-­-
BiOI heterojunction for degradation of crystal violet under simulated solar light irradiation.
Nano, 13, 1850070.
71. Lina, Y., Wua, S., Lia, X., Wua, X., Yanga, C., Zenga, G., Penga, Y., Zhoua, Q., & Lub,
L. (2018). Microstructure and performance of Z-scheme photocatalyst of silver phosphate
modified by MWCNTs and Cr-doped SrTiO3 for malachite green degradation. Applied
Catalysis B: Environmental, 227, 557–570.
72. Sharma, G., Gupta, V.  K., Agarwal, S., Kumar, A., Thakur, S., & Pathania, D. (2016).
Fabrication and characterization of Fe@MoPO nanoparticles: Ion exchange behavior
and photocatalytic activity against malachite green. Journal of Molecular Liquids, 219,
1137–1143.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 29

73. Akerdi, A. G., Bahrami, S. H., Arami, M., & Pajootan, E. (2016). Photocatalytic discoloration
of Acid Red 14 aqueous solution using titania nanoparticles immobilized on graphene oxide
fabricated plate. Chemosphere, 159, 293–299.
74. Mirzaei, A., Yerushalmi, L., Chen, Z., & Haghighat, F. (2018). Photocatalytic degradation of
sulfamethoxazole by hierarchical magnetic ZnO@g-C3N4: RSM optimization, kinetic study,
reaction pathway and toxicity evaluation. Journal of Hazardous Materials, 359, 516–526.
75. Ye, Y., Feng, Y., Bruning, H., Yntema, D., & Rijnaarts, H. H. M. (2018). Photocatalytic deg-
radation of metoprolol by TiO2 nanotube arrays and UV-LED: Effects of catalyst properties,
operational parameters, commonly present water constituents, and photo-induced reactive
species. Applied Catalysis B: Environmental, 220, 171–181.
76. Jallouli, N., Pastrana-Martínez, L. M., Ribeiro, A. R., Moreira, N. F. F., Faria, J. L., Hentati,
O., Silva, A. M. T., & Ksibi, M. (2018). Heterogeneous photocatalytic degradation of ibu-
profen in ultrapure water, municipal and pharmaceutical industry wastewaters using a TiO2/
UV-LED system. Chemical Engineering Journal, 334, 976–984.
77. Elhalil, A., Elmoubarki, R., Farnane, M., Machrouhi, A., Mahjoubi, F. Z., Sadiq, M., Qourzal,
S., & Barka, N. (2018). Photocatalytic degradation of caffeine as a model pharmaceutical
pollutant Mg doped ZnO-Al2O3 heterostructure. ENMM, 10, 63–72.
78. Sacco, O., Murcia, J. J., Lara, A. E., Hernandez-Laverde, M., Rojas, H., Naio, J. A., Hidalgo,
M. C., & Vaiano, V. (2020). Pt–TiO2–Nb2O5 heterojunction as effective photocatalyst for the
degradation of diclofenac and ketoprofen. Materials Science in Semiconductor Processing,
107, 104839.
79. Raizadaa, P., Kumaria, J., Dhiman, R., Singhb, V.  P., & Singha, P. (2017). Magnetically
retrievable Bi2WO6/Fe3O4 immobilized on graphene sand composite for investigation
of ­photocatalytic mineralization of oxytetracycline and ampicillin. Process Safety and
Environmental Protection, 106, 104–116.
80. Nguyen, T.  T., Nam, S.  N., Son, J., & Oh, J. (2019). Tungsten Trioxide (WO3)-assisted
photocatalytic degradation of amoxicillin by simulated solar irradiation. Scientific Reports,
9, 9349.
81. Tang, X., Wang, Z., & Wang, Y. (2018). Visible active N-doped TiO2/reduced graphene oxide
for the degradation of tetracycline hydrochloride. Chemical Physics Letters, 691, 408–414.
82. Hong, Y., Li, C., Yin, B., Li, D., Zhang, Z., Mao, B., Fan, W., Gu, W., & Shi, W. (2018).
Promoting visible-light-induced photocatalytic degradation of tetracycline by an efficient and
stable beta-Bi2O3@g-C3N4 core/shell nanocomposite. Chemical Engineering Journal, 338,
137–146.
83. Tahir, M.  B., Sagir, M., & Shahzad, K. (2019). Removal of acetylsalicylate and methyl-­
theobromine from aqueous environment using nano-photocatalyst WO3-TiO2 @g-C3N4 com-
posite. Journal of Hazardous Materials, 363, 205–213.
84. Akbarzadeh, R., Fung, C. S. L., Rather, R. A., & Lo, I. M. C. (2018). One-pot hydrothermal
synthesis of g-C3N4/Ag/AgCl/BiVO4 micro-flower composite for the visible light degradation
of ibuprofen. Chemical Engineering Journal, 341, 248–261.
85. Kumar, A., Sharma, S.  K., Sharma, G., Al-Muhtaseb, A.  H., Naushad, M., & Ghfar,
A. A. (2019). Wide spectral degradation of Norfloxacin by Ag@BiPO4/BiOBr/BiFeO3 nano-­
assembly: Elucidating the photocatalytic mechanism under different light sources. Journal of
Hazardous Materials, 364, 429–440.
86. Wanga, L., Yang, G., Wang, D., Lu, C., Guana, W., Li, Y., Deng, J., & Crittenden, J. (2019).
Fabrication of the flower-flake-like CuBi2O4/Bi2WO6 heterostructure as efficient visible-­
light driven photocatalysts: Performance, kinetics and mechanism insight. Applied Surface
Science, 495, 143521.
87. Khadgi, N., Li, Y., Upreti, A.  R., Zhang, C., Zhang, W., Wang, Y., & Wang, D. (2016).
Enhanced photocatalytic degradation of 17 α-ethinylestradiol exhibited by multifunctional
ZnFe2O4–Ag/rGO nanocomposite under visible light. Photochemistry and Photobiology, 92,
238–246.
30 Y. V. Divyasri et al.

88. Mi, X., Han, J., Sun, Y., Li, Y., Hu, W., & Zhan, S. (2019). Enhanced catalytic degradation by
using RGO-Ce/WO3 nanosheets modified CF as electro-Fenton cathode: Influence factors,
reaction mechanism and pathways. Journal of Hazardous Materials, 367, 365–374.
89. Verma, R., Singh, S., Dalai, M. K., Saravanan, M., Agrawal, V. V., & Srivastava, A. K. (2017).
Photocatalytic degradation of polypropylene film using TiO2-based nanomaterials under
solar irradiation. Materials and Design, 133, 10–18.
90. Kamalian, P., Khorasani, S.  N., Abdolmalek, A., & Neisiany, R.  E. (2019). Grafted ZnO
nanoparticles used for development in photocatalytic degradation performance of polyethyl-
ene. Polymer Bulletin, 76, 3593–3606.
91. Tu-morn, M., Pairoh, N., Sutapun, W., & Trongsatitkul, T. (2019). Effects of titanium diox-
ide nanoparticle on enhancing degradation of polylactic acid/low density polyethylene blend
films. Materials Today: Proceedings, 17, 2048–2061.
92. Soitong, T. (2018). Photo-degradation of polypropylene-ascorbic acid TiO2 composite films.
International Polymer Processing, 33, 29–34.
93. Koysuren, H.  N. (2018). Solid-phase photocatalytic degradation of polyvinyl borate.
Catalysts, 8, 499.
94. Tofa, T. S., Kunjali, K. L., Paul, S., & Dutta, J. (2019). Visible light photocatalytic degrada-
tion of microplastic residues with zinc oxide nanorods. Environmental Chemistry Letters, 17,
1341–1346.
95. Xu, F., Zhang, J., Zhu, B., Yu, J., & Xu, J. (2018). CuInS2 sensitized TiO2 hybrid nanofi-
bers for improved photocatalytic CO2 reduction. Applied Catalysis B: Environmental, 230,
194–202.
96. Zhang, S., Yin, X., & Zheng, Y. (2018). Enhanced photocatalytic reduction of CO2 to meth-
anol by ZnO nanoparticles deposited on ZnSe nanosheet. Chemical Physics Letters, 693,
170–175.
97. Bhosale, R., Jain, S., Vinod, C. P., Kumar, S., & Ogale, S. (2019). Direct Z-Scheme g-C3N4/
FeWO4 nanocomposite for enhanced and selective photocatalytic CO2 reduction under vis-
ible light. ACS Applied Materials & Interfaces, 11, 6174–6183.
98. Jiang, Z., Wan, W., Li, H., Yuan, S., Zhao, H., & Wong, P. K. (2018). A hierarchical Z-Scheme
α-Fe2O3/g-C3N4 hybrid for enhanced photocatalytic CO2 reduction. Advanced Materials, 30,
1706108.
99. Lingampalli, S. R., Ayyub, M. M., Magesh, G., & Rao, C. N. R. (2018). Photocatalytic reduc-
tion of CO2 by employing ZnO/Ag1-x Cux/CdS and related heterostructures. Chemical Physics
Letters, 691, 28–32.
100. Kim, C., Cho, K.  M., Al-Saggaf, A., Gereige, I., & Jung, H.  T. (2018). Z-scheme photo-
catalytic CO2 conversion on three-dimensional BiVO4/carbon-coated Cu2O nanowire arrays
under visible light. ACS Catalysis, 8, 4170–4177.
101. Y. W. Teh, Y. W. Goh, X. Y. Kong, B. J. Ng, S. T. Yong, S. P Chai, Fabrication of Bi2WO6/Cu/
WO3 all-solid-state Z-scheme composite photocatalyst to improve CO2 photoreduction under
visible light irradiation.
102. Tan, D., Zhang, J., Shi, J., Li, S., Zhang, B., Tan, X., Zhang, F., Liu, L., Shao, D., & Han,
B. (2018). Photocatalytic CO2 transformation to CH4 by Ag/Pd bimetals supported on
N-Doped TiO2 nanosheet. ACS Applied Materials & Interfaces, 10, 24516–24522.
103. Pan, Y.  X., You, Y., Xin, S., Li, Y., Fu, G., Cui, Z., Men, Y.  L., Cao, F.  F., Yu, S.  H., &
Goodenough, J.  B. (2017). Photocatalytic CO2 reduction by carbon-coated indium-oxide
nanobelts. Journal of the American Chemical Society, 139, 4123–4129.
104. Guan, J., Wang, H., Li, J., Maa, C., & Huo, P. (2019). Enhanced photocatalytic reduction of
CO2 by fabricating In2O3/CeO2/HATP hybrid multi-junction photocatalyst. Journal of the
Taiwan Institute of Chemical Engineers, 99, 93–103.
105. Jia, P., Guo, R., Pan, W., Huang, C., Tang, J., Liu, X., Qin, H., & Xu, Q. (2019). The MoS2/
TiO2 heterojunction composites with enhanced activity for CO2 photocatalytic reduction
under visible light irradiation. Colloids and Surfaces, 570, 306–316.
1  Nanostructures in Photocatalysis: Opportunities and Challenges for Environmental… 31

106. Cao, S., Shen, B., Tong, T., Fu, J., & Yu, J. (2018). 2D/2D Heterojunction of ultrathin MXene/
Bi2WO6 nanosheets for improved photocatalytic CO2 reduction. Advanced Functional
Materials, 28, 1800136.
107. Meng, A., Zhang, L., Cheng, B., & Yu, J. (2019). TiO2−MnOx−Pt, Hybrid multiheterojunc-
tion film photocatalyst with enhanced photocatalytic CO2-reduction activity. ACS Applied
Materials & Interfaces, 11, 5581–5589.
108. Xu, C., Qiu, P., Li, L., Chen, H., Jiang, F., & Wang, X. (2018). Bismuth subcarbonate with
designer defects for broad spectrum photocatalytic nitrogen fixation. ACS Applied Materials
& Interfaces, 10, 25321–25328.
109. Guo, L., Han, X., Zhang, K., Zhang, Y., Zhao, Q., Wang, D., & Fu, F. (2019). In-Situ con-
struction of 2D/2D ZnIn2S4/BiOCl heterostructure with enhanced photocatalytic activity for
N2 fixation and phenol degradation. Catalysts, 9, 729.
110. Zhang, C., Xu, Y., Lv, C., Bai, L., Liao, J., Zhai, Y., Zhang, H., & Chen, G. (2020). Amorphous
engineered cerium oxides photocatalyst for efficient nitrogen fixation. Applied Catalysis B:
Environmental, 264, 118416.
111. Liu, Q., Ai, L., & Jiang, J. (2018). MXene-derived TiO2@C/g-C3N4 heterojunctions for
highly efficient nitrogen photofixation. Journal of Materials Chemistry A, 6, 4102–4110.
112. Li, X., Sun, X., Zhang, L., Sun, S., & Wang, W. (2018). Efficient photocatalytic fixation of N2
by KOH treated g-C3N4. Journal of Materials Chemistry A, 6, 3005–3011.
113. Sun, S., An, Q., Wang, W., Zhang, L., Liu, J., & Goddard, W. A. (2017). Efficient photocata-
lytic reduction of dinitrogen to ammonia on bismuth monoxide quantum dots. Journal of
Materials Chemistry, 5, 201–209.
114. Wu, S., Tan, X., Liu, K., Lei, J., Wang, L., & Zhang, J. (2019). TiO2 (B) nanotubes with
ultrathin shell for highly efficient photocatalytic fixation of nitrogen. Catalysis Today, 335,
214–220.
115. Hirakawa, H., Hashimoto, M., Shiraishi, Y., & Hirai, T. (2017). Photocatalytic conversion of
nitrogen to ammonia with water on surface oxygen vacancies of titanium dioxide. Journal of
the American Chemical Society, 139, 10929–10936.
116. Li, Z., Gu, G., Hu, S., Zou, X., & Wu, G. (2019). Promotion of activation ability of N vacan-
cies to N2 molecules on sulfur-doped graphitic carbon nitride with outstanding photocatalytic
nitrogen fixation ability. Chinese Journal of Catalysis, 40, 1178–1186.
117. Wu, G., Yu, L., Liu, Y., Zhao, J., Han, Z., & Geng, G. (2019). One step synthesis of N vacancy-­
doped g-C3N4/Ag2CO3 heterojunction catalyst with outstanding “two-path” photocatalytic N2
fixation ability via in-situ self-sacrificial method. Applied Surface Science, 481, 649–660.
118. Nazemi, M., & El-Sayeda, M. A. (2019). Plasmon-enhanced photo(electro)chemical nitro-
gen fixation under ambient conditions using visible light responsive hybrid hollow Au-Ag2O
nanocages. Nano Energy, 63, 103886.
119. Liao, Y., Lin, J., Cui, B., Xie, G., & Hu, S. (2020). Well-dispersed ultrasmall ruthenium
on TiO2(P25) for effective photocatalytic N2 fixation in ambient condition. Journal of
Photochemistry and Photobiology A: Chemistry, 387, 112100.
120. Gao, X., An, L., Qu, D., Jiang, W., Chai, Y., Sun, S., Liu, X., & Sun, Z. (2019). Enhanced
photocatalytic N2 fixation by promoting N2 adsorption with a co-catalyst. Scientific Bulletin,
64, 918–925.
121. Zhang, G., Chen, D., Li, N., Xu, Q., Li, H., He, J., & Lu, J. (2018). SnS2/SnO2 heterostruc-
tured nanosheet arrays grown on carbon cloth for efficient photocatalytic reduction of Cr(VI).
Journal of Colloid and Interface Science, 514, 306–315.
122. Le, A. T., Pung, S. Y., Sreekantan, S., Matsuda, A., & Huynh, D. P. (2019). Mechanisms of
removal of heavy metal ions by ZnO particles. Heliyon, 5, e01440.
123. Dong, C., Lu, J., Qiu, B., Shen, B., Xing, M., & Zhang, J. (2018). Developing stretchable and
graphene-oxide-based hydrogel for the removal of organic pollutants and metal ions. Applied
Catalysis B: Environmental, 222, 146–156.
32 Y. V. Divyasri et al.

124. Zhang, H., Wang, X., Li, N., Xia, J., Meng, Q., Ding, J., & Lu, J. (2018). Synthesis and char-
acterization of TiO2/graphene oxide nanocomposites for photoreduction of heavy metal ions
in reverse osmosis concentrate. RSC Advances, 8, 34241.
125. Ghafoor, S., Hussain, S. Z., Waseem, S., & Arshad, S. N. (2018). Photo-reduction of heavy
metal ions and photo disinfection of pathogenic bacteria under simulated solar light using
photosensitized TiO2 nanofibers. RSC Advances, 8, 20354.
126. Du, Y., Zhang, S., Wang, J., Wu, J., & Dai, H. (2018). Nb2O5 nanowires in-situ grown on
carbon fiber: A high efficiency material for the photocatalytic reduction of Cr(VI). Journal of
Environmental Sciences, 66, 358–367.
127. Cui, Y., Li, M., Wang, H., Yang, C., Meng, S., & Chen, F. (2018). In-situ synthesis of sulfur
doped carbon nitride microsphere for outstanding visible light photocatalytic Cr(VI) reduc-
tion. Separation and Purification Technology, 199, 251–259.
128. Huang, Q., Liu, Y., Cai, T., & Xi, X. (2018). Simultaneous removal of heavy metal ions
and organic pollutant by BiOBr/ Ti3C2 nanocomposite. Advances in Colloid and Interface
Science, 254, 76–93.
129. She, P., Li, J., Bao, H., Xu, X., & Hong, Z. (2019). Green synthesis of Ag nanoparticles
decorated phosphorus doped g-C3N4 with enhanced visible-light-driven bactericidal activity.
Journal of Photochemistry and Photobiology A: Chemistry, 384, 112028.
130. Zhou, Q., Ma, S., & Zhan, S. (2018). Superior photocatalytic disinfection effect of Ag-3D
ordered mesoporous CeO2 under visible light. Applied Catalysis B: Environmental,
224, 27–37.
131. Liang, J., Deng, J., Liu, F., Li, M., & Tong, M. (2018). Enhanced bacterial disinfection by
Bi2MoO6-AgBr under visible light irradiation. Colloids and Surfaces B: Biointerfaces, 161,
528–536.
132. Zhang, B., Zou, S., Cai, R., Li, M., & He, Z. (2018). Highly-efficient photocatalytic disin-
fection of Escherichia coli under visible light using carbon supported vanadium tetrasulfide
nanocomposites. Applied Catalysis B: Environmental, 224, 383–393.
133. Xu, Y., Liu, Q., Liu, C., Zhai, Y., Xie, M., Huang, L., Xu, H., Li, H., & Jing, J. (2018).
Visible-­light-­driven Ag/AgBr/ZnFe2O4 composites with excellent photocatalytic activity
for E. coli disinfection and organic pollutant degradation. Journal of Colloid and Interface
Science, 512, 555–566.
134. Ma, S., Zhan, S., Xia, Y., Wang, P., Hou, Q., & Zhou, Q. (2019). Enhanced photocatalytic
bactericidal performance and mechanism with novel Ag/ZnO/g-C3N4 composite under vis-
ible light. Catalysis Today, 330, 179–188.
135. Yu, P., Zhou, X., Yan, Y., Li, Z., & Zheng, T. (2019). Enhanced visible-light-driven photocata-
lytic disinfection using AgBr modified g-C3N4 composite and its mechanism. Colloids and
Surfaces B: Biointerfaces, 179, 170–179.
136. Mathew, S., Ganguly, P., Rhatigan, S., Kumaravel, V., Byrne, C., Hinder, S. J., Bartlett, J.,
Nolan, M., & Pillai, S. C. (2018). Cu-Doped TiO2: Visible light assisted photocatalytic anti-
microbial activity. Applied Sciences, 8, 2067.
137. Ma, X., Xiang, Q., Lia, Y., Wen, T., & Zhang, H. (2018). Visible-light-driven CdSe quantum
dots/graphene/TiO2 nanosheets composite with excellent photocatalytic activity for E. coli
disinfection and organic pollutant degradation. Applied Surface Science, 457, 846–855.
138. Birben, N. C., Tomruk, A., & Bekbolet, M. (2017). The role of visible light active TiO2 speci-
mens on the solar photocatalytic disinfection of E. coli. Environmental Science and Pollution
Research, 24, 12618–12627.
139. Iqbal, T., Ali, F., Khalid, N. R., Tahir, M. B., & Ijaz, M. (2019). Facile synthesis and antimi-
crobial activity of CdS-Ag2S nanocomposites. Bioorganic Chemistry, 90, 103064.
140. Gao, Y., Mahmoudi, B., Fakhri, A., Aghazadeh, H., Hosseini, M., & Ebrahimi, H. A. (2019).
Synthesis of MnO2/CdTiO3 nano-structure for high performance photocatalysis and antimi-
crobial application. Applied Organometallic Chemistry, 33, e5051.
Chapter 2
Nanostructured Heterojunction (1D-0D
and 2D-0D) Photocatalysts
for Environmental Remediation

Lakshmana Reddy Nagappagari, Kiyoung Lee, Ajay Rakesh,


Subramanian Balakumar, and M. V. Shankar

2.1  Introduction

Environmental pollution from industries, automobiles, domestic usage, and sewage


activities has been constantly increasing day by day due to increasing population
and utilization of all these pollution-causing systems [1–3]. Hence these mentioned
human activities cause a major impact on water and air, which consequently damage
nature and affect human beings very severely. Therefore, there is an urgent need to
develop efficient technologies in a sustainable way to cop all these challenges and
make pollution-free environment for future generations. In this connection the
heterogeneous photocatalysis (PC) and photoelectrocatalysis (PEC) have become
emerging technologies for environmental applications [4–6]. Much attention is paid
especially on using various types of nanostructured photocatalysts due to their
unique nanoscale properties like high surface area, quantum confinement, and a
greater number of active sites for redox reactions on the surface of the photocatalysts
in the reaction medium [7, 8]. Various types of nanomaterials like 0D, 1D, 2D, and
3D nanostructures (Fig. 2.1) have well focused on the environmental remediation
such as organic dye degradation [9, 10], removal of metal ions [11–13], fluoride
removal [14–16], organic contaminants [17–19], and other environmental

L. R. Nagappagari · K. Lee
Department of Energy Chemical Engineering, School of Nano & Materials Science and
Engineering, Kyungpook National University, Sangju, Republic of Korea
A. Rakesh · S. Balakumar
National Centre for Nanoscience and Nanotechnology (NCNN), University of Madras,
Chennai, Tamil Nadu, India
M. V. Shankar (*)
Nanocatalysis and Solar Fuels Research Laboratory, Department of Materials Science
& Nanotechnology, Yogi Vemana University, Kadapa, Andhra Pradesh, India
e-mail: shankar@yogivemanauniversity.ac.in

© Springer Nature Switzerland AG 2021 33


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_2
34 L. R. Nagappagari et al.

Heterogeneous Nanostructured Materials with Different Morphologies


0-D 1-D 2-D 3-D
g) j)
a) d)

Carbon Nanotube
Core-Shell Nanoparticle
Based Composite Electrode Graphene Based Composite Mesoporous Composite Electrode
b) e) h) k)

Nanoparticles Encapsulated
Coaxial Nanowire Array Carbon Coated Nanoplates Microporous Composite Electrode
in Hollow Nanosphere
c) f) i) l)

Composite Nanoparticle Composite Nanowire Array Carbon Coated Nanobelts Future 3-D Electrode

Fig. 2.1  Different types of nanostructured materials and morphologies used for heterogeneous
photocatalytic environmental remediation: (a–c) 0D nanostructures, (d–f) 1D nanostructures, (g–i)
2D nanostructures, and and (j–l) 3D nanostructures. Reprinted with permission from Ref. [23],
Copyright @ Royal Society of Chemistry 2011

applications. More importantly, the heterojunction nanostructures like 1D-0D and


2D-0D heterojunction photocatalysts have acquired important attention [20–22],
since the heterojunction formed between the interface of two materials will
effectively minimize the charge carrier recombination and thus improve the catalytic
activity.

2.2  Impact of Nanostructures on Material Properties

More importantly, the nanostructures will obviously have a major impact on mate-
rial properties, for example, 0D quantum dots show surface plasmon resonance
(SPR) effect to improve the catalytic activity by absorption of visible photons [24,
25]; the 1D nanostructures like nanotubes, nanorods, and nanobelts show high sur-
face area, quantum confinement along the axis of the nanostructures, and unidimen-
sional flow of electrons [10, 26, 27]; and the 2D nanosheets show high surface area
and more number of active sites to improve the photocatalytic efficiency [28–30].
Hence the study of these types of nanostructured material properties for environ-
mental application has recorded as an important task. Figure 2.2 shows the impor-
tant properties of nanostructures especially the properties of the material useful for
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 35

Bulk semiconductor Nanocrystal semiconductor Molecule

Energy
Confinement width
Bandgap Eg

Decreasing nanocrystal size

Quantum
confinement

Properties of
Nanostructures

Defect site Active site

High surface area

Fig. 2.2  Important properties of nanostructures highly useful for environmental remediation

environmental remediation. Therefore, in this book chapter, we focus on a detailed


discussion of various types of heterojunction nanostructured photocatalysts for
environmental applications.

2.3  1 D-0D Heterojunction Photocatalysts


for Pollutant Degradation

Heterojunction photocatalysts have become the hot topic of research for environ-
mental photocatalysis, due to improved efficiencies acquired by minimized charge
carrier recombination, the interface of the heterojunction [31, 32]. Therefore, a wide
variety of heterojunction photocatalysts, viz., p-n junction, n-p junction, type I het-
erojunction, type II heterojunction, Schottky junction, and Z-scheme heterojunc-
tion, have been investigated for environmental remediation [33–39]. Figure  2.3
36 L. R. Nagappagari et al.

Fig. 2.3  Various types of heterojunctions formed by semiconductor photocatalysts for environ-
mental applications. (a) Type I heterojunction, (b) type II heterojunction, (c) p-n junction, (d)
Schottky junction, and (e) Z-scheme heterojunction. (Note: A, D, and Ef represent electron accep-
tor, electron donor, and Fermi level, respectively.) Reprinted with permission from Ref. [40].
Copyright @ Springer Science 2019

displays the various types of heterojunction photocatalysts used for environmental


remediation [40]. More importantly, 1D-0D and 2D-0D heterojunction photocata-
lysts have become the important materials for various pollutants’ degradation.
Hence, we mainly focus on detailed discussion on this important field of research in
this book chapter.

2.3.1  M
 etal Oxide-Metal Oxide (1D-0D) Heterojunction
Photocatalysts for Pollutant Degradation

1D-0D heterojunction photocatalysts consisting of metal oxide-metal oxide


(MO-MO)-based materials will occupy the top place. Unlike insulators or metals,
semiconductor metal oxides have the bandgap that generates electron(e−)/hole (h+)
pairs when they absorb a photon with energy equal to their bandgap or higher than
that. Semiconducting metal oxides are commonly used as photocatalysts due to
their favorable bandgap (Eg) and band edge positions. The controlled morphological
and textural features, variable surface chemistry, high surface area, specific
crystalline nature, and abundant availability make the nanostructured metal oxides
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 37

and their composites highly selective materials for efficient removal of organic
pollutants based on adsorption and photocatalytic degradation [41]. More
importantly, the metal oxides show good stability and a greater number of active
sites. Especially the 1D TiO2 nanostructure-based heterojunctions with other 0D
metal oxides (MnO2, CuxO, Ag2O, ZnO, etc.) have been well studied for various
pollutants’ degradation, since, as a photocatalyst, titanium dioxide (TiO2) has
attracted substantial attention for a long time and is considered as one of the most
promising materials for commercial use due to its outstanding optical and electronic
properties, photoactivity, high chemical stability, low cost, non-toxicity, reusability,
and eco-friendliness [17, 42–44]. Therefore, modification of 1D TiO2 with other
MO photocatalysts will tremendously improve the catalytic efficiencies. The
advantages of these types of photocatalysts will be discussed in detail as follows.

TiO2-MnO2 Heterojunction

Recent studies have focused on TiO2-MnO2 system due to MnO2’s non-toxicity and
earth abundance, despite the narrow bandgap of MnO2 (0.26–2.7 eV), which could
allow the absorption of visible and theoretically even infrared light [45] which has
attributed to the improved photocatalytic efficiency and effective separation of
photogenerated electrons and holes. Martinez et al. [46] studied the self-organized
TiO2–MnO2 nanotube arrays for efficient photocatalytic degradation of toluene. The
excitation mechanism of TiO2−MnO2 NTs under visible light was presented,
pointing out the importance of MnO2 species for the generation of e− and h+ pairs.
Crystal structure and self-organized TiO2−MnO2 nanotube arrays were successfully
obtained by one-step anodic oxidation of Ti-Mn alloys in an ethylene glycol-based
electrolyte as shown in Fig. 2.4. In this composite photocatalyst, the MnO2 treated

Fig. 2.4  (a) XRD spectra of pristine TiO2 and TiO2−MnO2 NTs elucidating the effect of anodiza-
tion potential, (b) top-view and cross-sectional SEM images of pristine TiO2 and TiO2–MnO2 NTs
(the effect of applied voltage, manganese content in the Mn/Ti alloy, and water content in the
electrolyte on the morphology of formed nanotubes) and EDX mapping of the Ti90Mn10_30V sam-
ple. Reprinted with permission from ref.[46] Copyright @ MDPI 2017
38 L. R. Nagappagari et al.

as co-catalysts and photogenerated holes from VB of MnO2 could be involved in the


formation of hydroxyl radicals (˙OH) as part of toluene degradation mechanism.

TiO2-CuO Heterojunction

TiO2–CuO heterojunction photocatalysts are also one of the promising candidates


for various environmental applications. Due to the combination of the wide bandgap
and narrow bandgap semiconductors, this type of heterojunction allows the
maximum utilization of solar spectrum [47]. Lee et al. [48] reported the structured
electrospun TiO2/CuO composite nanofibers for high efficient photocatalytic
cogeneration of clean water and energy from dye wastewater. The TiO2/CuO
composite nanofibers demonstrated multifunctional ability for concurrent
photocatalytic organic degradation and H2 generation from dye wastewater. The
enhanced photocatalytic activity of TiO2/CuO composite nanofibers was ascribed to
its excellent synergy of physicochemical properties: (1) mesoporosity and large
specific surface area for efficient substrate adsorption, mass transfer, and light
harvesting; (2) redshift of the absorbance spectra for enhanced light utilization; (3)
nanofibrous structure for efficient charge transfer and ease of recovery; (4) TiO2/
CuO heterojunctions which enhance the separation of electrons and holes; and (5)
presence of CuO which serves as co-catalyst for the H2 production. The photocatalytic
activity and reaction mechanism of TiO2/CuO for AO7 removal by adsorption from
wastewater was shown in Fig. 2.5.

TiO2–Ag2O Heterojunction

Silver-based oxides are the most suitable for industrial application because of its
high efficiency, low cost, and easy preparation. Ag2O has acquired more importance
especially in environmental and biological applications such as dye degradation and
antibacterial disinfection [49, 50]. Ag2O is a p-type semiconductor with an energy
bandgap of 1.46 eV [51]. All these properties of Ag2O are beneficial for the formation
of p-n nanoheterojunction with TiO2 for superior photocatalyst. The photogenerated
electrons can move to the conduction band of n-type TiO2, and photogenerated
holes can move to the valence band of p-type Ag2O which promotes an interfacial
electron transfer process and reduces the charge recombination on the semiconductor
and eventually improves the degradation efficiency. Sarkar et al. [52] reported three-­
dimensional Ag2O/TiO2 type II (p-n) nanoheterojunctions for superior photocatalytic
activity toward the degradation of methyl orange (MO) aqueous solution. The
prepared heterojunction photocatalyst with the optimum molar ratio of TiO2
andAgNO3 (4:1) degrades the methyl orange (MO) within 15 min of UV irradiation
which is three times faster than that of the pure TiO2. Figure 2.6 shows the photo-
catalytic degradation profiles of Ag2O/TiO2 (p-n) nanoheterojunctions.
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 39

Fig. 2.5  (a) Removal of AO7 by PCO at t = 3 min during the UV–visible irradiation on the pho-
tocatalyst suspension, (b) schematic diagram of the photogenerated electron and hole transfer
between the TiO2/CuO heterojunctions. Reprinted with permission from ref. [48] Copyright @
Elsevier 2013

TiO2–ZnO Heterojunction

As we all know, titanium dioxide (TiO2) and zinc oxide (ZnO) are two important
semiconductors for photocatalytic applications widely studied for dye degrada-
tion, wastewater treatment, and environmental pollution treatments [10, 27, 53,
54]. Therefore, this important combination of materials can obviously make rev-
olutionary changes in photocatalysis. The positions of VB and CB of TiO2 are
2.94 and −0.30 eV, and the VB and CB positions of ZnO are 2.88 and −0.32 eV,
respectively [55]. It is clear that the positions of VB and CB for TiO2 and ZnO are
close to each other [56]. In theory, a good heterojunction photocatalyst will be
formed when the two semiconductors are combined. Sun et al. [57] investigated
the photocatalytic activities of the ZnO/TiO2 photocatalysts for degradation of
rhodamine B (RhB), methyl orange (MO), and bisphenol A (BPA) under UV
light illumination. The transfer mechanisms of the photoexcited carriers for the
ZnO/TiO2photocatalysts were proposed and verified. Figure 2.7 displays the ESR
signals of 5,5-dimethyl-1-­pyrroline-N-oxide (DMPO) for identification of ˙O2¯
40 L. R. Nagappagari et al.

Fig. 2.6  (a) Photocatalytic degradation profiles of MO under UV irradiation. (b) Kinetic plot with
the different molar ratios of TiO2 to AgNO3. (c) Degradation percentages after 15- and 30-min
irradiation. (d) Digital photograph of the decolorization of MO after different UV exposure times
for TiO2/AgNO3 = 4:1. (e) 3D heterostructure after the photocatalytic performance. Reprinted with
permission from ref.[52] Copyright @ American Chemical Society 2013

and ˙OH radicals generated in the photocatalytic process and photoexcited elec-
tron-hole transfer process over the p-ZnO/n-TiO2.

2.3.2  M
 etal Oxide-Metal (1D-0D) Heterojunction
Photocatalysts for Pollutant Degradation

1D metal oxides-0D metals (MO-M) making as 1D-0D heterojunction photocata-


lysts have been considered as another important class of photocatalytic materials
widely useful for environmental remediation and organic dye degradation applica-
tions. More importantly, TiO2-M (M = Ag, Au, Pt, Cu, Fe, Cr, etc.)-based nanostruc-
tured materials are extensively studied for wastewater treatment, organic dye
degradation, and other environmental applications. Because metal nanoparticles
decorated on TiO2 nanostructures play an important role (either co-catalyst of sen-
sitizer role) in the catalytic process, they can effectively capture the charge carriers
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 41

a b DMPO- • OH
DMPO- • O2- ZnO ZnO

99% ZnO/TiO2
99% ZnO/TiO2
Intensity (a.u)

95% ZnO/TiO2

Intensity (a.u)
95% ZnO/TiO2

5% ZnO/TiO2
5% ZnO/TiO2
1% ZnO/TiO2
1% ZnO/TiO2
TiO 2 TiO2

Blank Blank

3480 3500 3520 3540 3480 3500 3520 3540


Magnetic field (G) Magnetic field (G)

c d e
O2 • O2
Evac e- e- e- e- e- e-
-

e- e- e- e- e- e-
Eg = 3.24 eV

4.38 eV

Eg = 3.24 eV
4.85 eV

Eg = 3.2 eV
Eg = 3.2 eV
Re
co

CB
m
bin

CB
Ef
e

H2O
Ef
VB
VB h+ h+ h+ h+ h+ h+
h+ h+ h+ h+ h+ h+
TiO 2 Zno n- TiO 2 p- Zno
• OH p-TiO 2 n-Zno

Fig. 2.7  ESR signals of 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) for identification of (a) ˙O2¯
and (b) ˙OH radicals generated in the photocatalytic process and (c) Energy-level line-up diagrams
for the (103) plane of TiO2 and the (100) plane of ZnO. (d) Schematic diagrams of the photoexcited
electron-hole transfer process over the p-ZnO/n-TiO2 (the main part of the ZnO/TiO2 composite is
TiO2) and (e) p-TiO2/n-ZnO (the main part of the TiO2/ZnO composite is ZnO) heterojunctions.
(CB, VB, e−, h+, Evac, and EF are conduction band, valance band, photogenerated electrons, holes,
vacuum level, and Fermi level, respectively.) Reprinted with permission from ref. [57] Copyright
@ American Chemical Society 2018

for enhancement of degradation efficiency [58, 59]. Therefore, in this section, a


detailed study has been investigated on TiO2-metal nanostructured photocatalysts.

TiO2–Ag Heterojunction

TiO2-decorated Ag nanostructures play an important role in dye degradation due to


surface plasmon resonance (SPR) effect and nanoscale properties of Ag nanoparticles
(NPs). Fei et al. [60] synthesized porous TiO2 nanowire-decorated Ag nanoparticles
by sol–gel method. The morphology of the 1D nanostructures was confirmed by
SEM and TEM analysis, and the schematic representation of the formation of 1D
nanotrusses was shown in Fig.  2.8. The prepared materials performed 98%
42 L. R. Nagappagari et al.

(a) e Sol-Gel Calcination


tur
e ra
temp
low
Melamine
AgNO 3 Calcination
Sol-Gel
hig
h
te
m
pe
ra
tu
re

AgNO 3-Melamine AgNO 3-Melamine -Ti(OH)4 TiO2-Ag

(b)

Fig. 2.8  (a) Schematic illustration of the controlled fabrication of porous TiO2−Ag nanostructures
templating supramolecular assembly with different morphologies. (b) TEM image of TiO2–Ag
nanotubes. Reprinted with permission from ref. [60] Copyright @ Wiley VCH 2014

degradation efficiency of MB dye which is much higher compared to commercial


TiO2 P25-Ag materials which showed only 78% under the same experimental
condition. Thus, the effect of 1D nanostructures is clearly beneficial for improved
catalytic efficiency.

TiO2–Au Heterojunction

TiO2–Au (1D-0D) nanostructures have also become an important combination of


photocatalysts for environmental pollutant degradation. Especially SPR effect of Au
NPs and 1D morphology of TiO2 shows the synergistic effect for many energy and
environmental applications [61–63]. Wen et al. [61] reported Au-decorated anatase/
rutile mixed TiO2 nanotubes for dye degradation. The optimal visible photocatalytic
activity was found in the sample Au3(DP350)/TiO2 with a loading of 3 wt% Au NPs
and calcining at 350  °C.  Through transmission electron microscopy, Au NPs of
4.16  nm diameter were observed at the interface between the anatase and rutile
phases in the optimal Au3(DP350)/TiO2 sample, and these joint active sites at the
interface were beneficial for charge separation. Later on Yu et al. [64] reported UV
and visible light photocatalytic activity of Au/TiO2 branched nanowires with
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 43

Fig. 2.9  A schematic diagram showing the formation of the (a) branched TiO2 nanowires
(Branched-NW), (b) those after the Au loading (Branched-NW-Au), and (c) the branched TiO2
nanowires with an intermediate Au loading (Branched-Au-NW). Reprinted with permission from
ref. [64] Copyright @ Nature 2017

anatase/rutile phase junctions and controlled Au locations. Figure  2.9 shows the
schematic diagram for the formation of the branched TiO2 nanowires decorated with
Au NPs. In this work, the Au decoration increased the photocatalytic activity under
the illumination of either UV or visible light, because of the beneficial effects of
either electron trapping or localized surface plasmon resonance (LSPR). Recently
Li and co-workers studied the TiO2 and Au–TiO2 nanomaterials for rapid photocata-
lytic degradation of antibiotic residues in aquaculture wastewater [65]. Therefore,
TiO2/Au nanostructures have been recognized as one of the important photocata-
lysts for environmental applications.

TiO2-Pt Heterojunction

Another important combination of photocatalysts will be TiO2–Pt heterojunctions,


since the Pt is well recognized as the best co-catalyst on the TiO2 support for
enhanced photocatalytic efficiency. Therefore, many TiO2–Pt-based nanostructures
have been investigated for environmental applications. For example, TiO2
nanotubes-Pt NPs, TiO2 nanorods-Pt NPs, and TiO2 nanofibers-Pt NPs have been
prepared by different methods like hydrothermal method, wet impregnation,
44 L. R. Nagappagari et al.

Fig. 2.10  (a) Photographs of photocatalytic degradation of MO dye with TNT, PTP-1.0, and
PTC-1.0 under solar light irradiation and (b) plausible reaction mechanism in Pt-TiO2 nanotube
photocatalyst prepared by photodeposition and chemical reduction methods. Reprinted with
permission from ref. [10]. Copyright @ Elsevier 2019

chemical reduction, and anodization and applied for energy and organic dye
degradation [66–68]. Recently in our previous reports, we investigated TiO2 nano-
tubes decorated by Pt quantum dots and prepared by chemical reduction and photo-
deposition methods and studied the effect of synthesis methods on valence states of
Pt that has readily influenced the solar light degradation efficiency [10].
Figure 2.10(a) displays the photographic images of photocatalytic degradation of
methyl orange (MO) dye for pristine TiO2 nanotubes (TNT), Pt-TNT, prepared by
chemical reduction and photodeposition methods. Here the Pt deposited on TNT by
chemical reduction method (PTC-1.0) displayed higher efficiency compared to one
prepared by photodeposition (PTP-1.0) and pristine TNT.  This is due to particle
reduction of Pt which consists of Pt2+ and Pt4+, as evidenced by XPS results in the
study [10]. Figure 2.10(b) shows the photocatalytic mechanism for MO degradation
and hydrogen generation of Pt-TNT nanomaterials prepared by photodeposition
and chemical reduction methods. Therefore, it was proved that not only the particle
size of Pt can influence the photocatalytic efficiency but also the valence states of Pt
can highly affect the degradation activities.

TiO2–Cu Heterojunction

TiO2 nanotube-decorated Cu nanoparticles can also be recognized as efficient pho-


tocatalysts for energy and environmental applications [69–72], because Cu NPs can
also act as very good co-catalyst and able to form heterojunctions [73–75]. Suhaimy
et al. [76] studied the nanosized copper (Cu)-incorporated TiO2 nanotubes synthe-
sized via the anodic oxidation technique in ethylene glycol (EG) containing 0.5 wt
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 45

% NH4F and 1.6  wt% KOH for the photocatalytic degradation of simazine
(2-chloro-4, 6-diethylamino-1,3,5-triazine) under ultraviolet (UV) illumination.
Influence of different loading Cu concentrations on the formation of Cu–TiO2 nano-
tube film was studied. The improved activity was claimed due to uniform surface
covering of the Cu-loaded TiO2NTs which acted as electron traps, preventing the
recombination of electron hole pairs, eventually leading to higher photocatalytic
activity.

2.4  2 D-0D Heterojunction Photocatalysts


for Pollutant Degradation

Recently, 2D nanomaterials have become the fascinating photocatalysts for many


catalytic applications including organic dye degradation, environmental air
purification, toxic gas sensing, etc. [28, 77, 78]. Especially the heterojunctions
made with graphitic carbon nitride (g-C3N4) nanosheets, graphene oxide (GO),
reduced graphene oxide (rGO), and carbon nanosheets have become the most
important materials [29, 79–81]. Due to their unique physicochemical properties,
2D material-based photocatalysts are expected to offer intriguing features such as
porous structures, high specific surface areas, good crystallinity, rich options of
host-guest species, better charge carrier separation, and abundant surface active
sites [82–84]. Hence in this section we mainly focus on heterojunctions of graphitic
carbon nitride (g-C3N4) nanosheets, graphene oxide (GO), reduced graphene oxide
(rGO), and carbon nanosheets for environmental applications. The structures of
these 2D nanosheets can be seen in Fig. 2.11.

2.4.1  2 D Nanosheet-Metal Oxide (2D-0D) Heterojunction


Photocatalysts for Pollutant Degradation

The 2D nanosheets and metal oxide (2D-0D) heterojunction photocatalysts are


important materials due to the special properties acquired by the nanosheets like
high electron transport properties, high surface area, and the efficient oxidation and
reduction properties of metal oxides after making heterojunction with the nanosheets
[85–87]. Therefore, in this part, we aim to focus on nanosheets coupled with metal
oxides as an efficient environmental photocatalyst. Clearly, the g-C3N4/metal oxide
(g-C3N4/MO), graphene oxide/metal oxide (GO/MO), reduced graphene oxide/
metal oxide (rGO/MO) heterojunction photocatalysts will be discussed.
46 L. R. Nagappagari et al.

Fig. 2.11  Important structures of 2D nanomaterials g-C3N4, graphene, and graphene oxide

g-C3N4/MO Heterojunction

The metal oxides have become one of the most important strategies to modify with
g-C3N4 for photocatalytic applications. Exploration of two different semiconductor
photocatalysts (nanosheets-metal oxides) with suitable valence band and conduc-
tion band potentials is considered to be one of the most effective strategies for accel-
erating charge transfer and separation by taking full consideration with the space
charge exhaustion and accumulation between two semiconductors. Till date, many
researchers have reported various metal oxide semiconductor/g-C3N4 heterojunc-
tions for photocatalytic environmental applications. It is reported that TiO2, ZnO,
BiVO4, SnO2, In2O3, etc. are promising candidates to form heterojunctions with
g-C3N4 [88–91]. In general, integrating TiO2 with g-C3N4 is considered to be the
most feasible strategy to expand its spectral scope to visible light as well as to
restrain the recombination process, thus resulting in improved performance. Various
TiO2 nanostructures, such as 0D nanoparticles, 1D nanowires, 2D nanosheets, and
3D mesoporous crystals, can be readily loaded on the g-C3N4 surface as an anchor-
ing site by a seed-induced solvothermal treatment [92]. It is reported that well-dis-
persed In2O3 nanocrystals on sheetlike g-C3N4 surfaces form intimate contact, which
allows for effective interfacial charge transfer across the In2O3/g-C3N4 heterojunc-
tion since the conduction band (CB) and valance band (VB) positions of In2O3
(−0.6  eV and +2.2  eV, respectively, vs. NHE) [93] are both lower than those of
g-C3N4 (−1.1 eV and + 1.6 eV, respectively, vs. NHE) [94]. In particular, the con-
duction band edge (−0.6 eV) of In2O3 is high enough for reduction of proton and
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 47

CO2. Cao et al. [88] studied the In2O3/g-C3N4 heterojunction for photocatalytic H2
generation and CO2 reduction as shown in Fig. 2.12; here the In2O3/g-C3N4 hetero-
junction performed the highest photocurrent response compared to corresponding
pristine materials. The reduced intensities in PL and time-resolved PL spectra also
confirm the improved charge carrier recombination and higher efficiencies. Finally,
the charge transfer mechanism in In2O3/g-C3N4 heterojunction was shown in
Fig. 2.12(d).

GO/MO Heterojunction

Graphene or graphene oxide (GO) has also been widely used as a support to enhance
the photocatalytic properties of the semiconductor metal oxide for the degradation
of organic species, wastewater treatments, and other environmental applications
[83, 84, 95]. The graphene-based lattice and existence of various oxygen-containing
groups (mainly epoxy and hydroxyl groups) enable GO abundant fascinating

Fig. 2.12  (a) Transient photocurrent responses for the g-C3N4, In2O3, and 10  wt%
In2O3/g-C3N4samples, (b) steady-state PL spectra and (c) time-resolved transient PL decay for pure
g-C3N4 and the 10 wt% In2O3/g-C3N4, and (d) schematic illustration of the photocatalytic process
for H2 evolution and CO2 reduction on the In2O3/g-C3N4 nanohybrids. Reprinted with permission
from Ref. [88] Copyright @ Elsevier 2014
48 L. R. Nagappagari et al.

properties [83]. First, the functional groups on GO surface act as effective anchoring
sites to immobilize various active species. Furthermore, GO possesses tuneable
electronic properties. In this connection, the heterojunction of GO with various
metal oxides like ZnO, In2O3, MoO3, V2O5, TiO2, BiVO4,etc. is a promising candidate
for environmental applications with improved efficiencies [96–99]. Rakkesh et al.
[97] reported nanostructuring of a (graphene nanosheets, GNS) GNS–V2O5–TiO2
core-shell photocatalyst for water remediation applications under sunlight
irradiation. The sunlight-active photocatalytic properties of the GNS–V2O5–TiO2
nanoarchitectures have been evaluated by photodegradation of acridine orange
(AO) dye in an aqueous medium. Photocatalytic degradation results showed 40%
for pure V2O5 nanorods within 60  min, 95% for V2O5–TiO2 core-shell nanorods
within 60  min, and 100% for GNS–V2O5–TiO2 nanoarchitectures within 20  min.
Thus, GNS played an important role in charge carrier transport and minimization of
recombination rate. Another study reported the role of the interfacial charge transfer
process in the graphene–ZnO–MoO3 core-shell nanoassemblies for efficient disin-
fection of industrial effluents [100]. In this study, a combined wet chemical strategy
was adopted to fabricate size controllable ZnO–MoO3 core-shell nanostructures by
varying the surface potential in the reaction medium. The layered MoO3 was
adsorbed on the surface of ZnO particles by electrostatic interaction and simultane-
ously anchored onto graphene nanosheets (GNS) by chemical bonds. Figure 2.13
shows the schematic representation of the graphene-ZnO–MoO3 core-­shell nanoas-
semblies and corresponding TEM images.

rGO/MO Heterojunction

Besides GO, the reduced graphene oxide (rGO) has also been widely used as sup-
port to enhance the photocatalytic properties of the semiconductor metal oxide for
the degradation of organic species [101–103]. Generally, GO is used to prepare the
rGO using a variety of reduction methods such as green, chemical, thermal, electro-
chemical, solvothermal, and photodeposition techniques [103–105]. Different types
of binary and ternary composites of rGO with large numbers of metal oxides such
as ZnO, CuO, TiO2, ZnFe2O4, ZrO2, and so forth have been reported as the active
photocatalyst for the degradation of organic compounds [106–109]. Mady et  al.
[106] reported facile microwave-assisted green synthesis of Ag-ZnFe2O4@rGO
nanocomposites for efficient removal of organic dyes under UV and visible light
irradiation. The heterojunction formed between rGO and ZnFe2O4 helped the fast
transfer of charge carriers for photocatalytic reactions and accordingly improved the
efficiency. In another study Wang et al. [99] studied the electrostatic self-assembly
of BiVO4-rGO nanocomposites for highly efficient visible light photocatalytic
activities. In this report, the smaller particle size with high surface areas and
increased interfacial interaction in BiVO4-rGO leads to increased photocatalytic
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 49

Fig. 2.13  (a) Schematically represented the synthesis procedure of GNS–ZnO–MoO3 hybrid
nanoassemblies by wet chemical techniques, (b) TEM images of GNS–ZnO–MoO3, and (c)
magnified image of the GNS–ZnO–MoO3 hybrid nanoassemblies. Remodified with permission
from Ref. [100]

reaction sites, extended photoresponding range, and enhanced photogenerated


charge separation and transportation efficiency. Very recently Mandal and co-work-
ers synthesized ZnO quantum dot-decorated rGO by soft chemical and solvother-
mal methods and successfully treated for degradation of methylene blue (MB) and
rhodamine 6G as shown in Fig. 2.14. In this work, the composite with ZnO and rGO
in 3:1 wt ratio degraded almost 100% of dyes under UV irradiation within 15, 25,
and 30 min, showing good efficacy in dye degradation. Photoluminescence lifetime
measurement revealed the charge transfer process from rGO to ZnO resulting in a
prompt recombination of the UV light-induced electron-­hole pair in ZnO/rGO
composite.
50 L. R. Nagappagari et al.

Fig. 2.14  (a) Schematic representation of the formation mechanism of ZnO/rGO composite, (b)
mechanism of the photodegradation of dye using ZnO/rGO composites as a catalyst. Adapted with
permission from Ref. [109] Copyright Elsevier 2019

2.4.2  2 D Nanosheet-Metal (2D-0D) Heterojunction


Photocatalysts for Pollutant Degradation

Metal nanoparticles perform special properties when combining with 2D support


materials. Especially metal nanoparticles consist of high surface to volume ratio,
high functionality, abundant active sites, and surface plasmon resonance (SPR)
effect in case of noble metals that are highly useful for photocatalytic redox reactions
when forming a Schottky junction [17, 25, 110–114]. Hence the heterojunction
photocatalysts made with 2D nanosheet-metal NPs are widely used in environmental
applications such as Ag, Au NPs-g-C3N4 [115], Pt-rGO, rGO–TiO2–Ag [116],
GO-Ag [117], and Pd–Ag–GO [118]. Hence in this section more detailed discussion
will be carried out as mentioned photocatalytic materials.
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 51

Fig. 2.15  (a) ESR signals of DMPO-OH adducts produced by Au/CNNS/W18O49, CNNS/W18O49,
CNNS, and W18O49, respectively. (b) PL spectra of Au/CNNS/W18O49, CNNS/W18O49, and
CNNS. (c) Transient photocurrent response of Au/CNNS/W18O49, CNNS/W18O49, CNNS, and
W18O49. (d and e) Schematic illustration of the proposed mechanism for Cr(vi) photoreduction:
switching charge transfer from (d) type II in CNNS/W18O49 to (e) direct Z-scheme in Au/CNNS/
W18O49. Reprinted with permission from Ref. [119], Copyright @ Royal Society of Chemistry 2018

Shi et al. [119] investigated the charge transfer mechanism by switching from
type II to Z-scheme heterojunction by deposition of Au NPs on g-C3N4 nanosheets
(CNNS). As a result, the designed Au/CNNS/W18O49 heterostructure shows
enhanced photocatalytic performance for Cr(VI) reduction than the CNNS/W18O49
heterostructure and single components (CNNS and W18O49) under visible light
irradiation. The ESR signals to detect ˙OH, PL spectra, transient photocurrent
density of pristine material and heterostructures of Au/CNNS/W18O49 and its charge
transfer mechanism are shown in Fig. 2.15. In summary, this study unveils that the
decoration of Au NPs onto CNNS can switch the interface charge transfer
heterojunction routes of CNNS/W18O49 from conventional type II to direct Z-scheme.
Contributions made by the synergistic interaction of the “electron sink” combined
with the Z-scheme heterojunction charge transfer mechanism allowed for the
enhanced charge transfer and separation efficiency of charge carriers by the ­Au/
CNNS/W18O49 composite that ultimately exhibited significant improvements in
photocatalytic activity for Cr(vi) reduction.
Very recently Satish et al. [120] studied the real-time photodegradation of meth-
ylene blue (MB), an organic pollutant, in the presence of sunlight at ambient tem-
perature using platinum-decorated reduced graphene oxide (rGO/Pt) nanocomposite.
The photocatalyst was prepared via a simple, one-pot, and green approach with the
52 L. R. Nagappagari et al.

Fig. 2.16  (a) Schematic of the synthesis of rGO/Pt nanocomposite using aqueous honey with GO
and Pt salt as precursor materials. (b) UV–vis absorption spectra of the rGO/Pt4 nanocomposite,
GO and H2PtCl6 salt; the inset figures show photographs of (I) GO, (II) H2PtCl6 salt, and (III) rGO/
Pt nanocomposite. (c) FTIR spectra of the synthesized rGO/Pt4 nanocomposite and GO. Reprinted
with permission from Ref. [120] Copyright @ Royal Society of Chemistry 2020

simultaneous reduction of GO and Pt using aqueous honey as a reducing agent.


Moreover, the honey not only simultaneously reduced Pt ions and GO but also
played a key role in the growth and dispersion of Pt nanoparticles on the surface of
rGO. The high percentage of Pt nanoparticles with an average size of 2.5 nm dis-
persed on rGO has shown excellent electrochemical performance. Figure 2.16 dis-
played the synthesis procedure of rGO/Pt nanocomposite, UV–vis absorption
spectra, and FTIR spectra of rGO and rGO/Pt.
Besides GO-Ag NPs also performed antibacterial activity [121] and elemental
mercury (Hgο) removal [122]. Rajesh et al. [123] studied the stabilization of silver
and gold nanoparticles (Ag/Au NPs) on graphene oxide (GO) functionalized with
PAMAM dendrimers. We investigated the photocatalytic activity toward degradation
of ozo dyes, namely, methyl orange and Congo red within a few seconds.
Figure 2.17(a, c) shows the catalytic degradation of MO with time, and it can be
observed that complete degradation occurs within 80 s and 150 s in the presence of
the Ag and Au NPs/GO-G3PAMAM nanocatalysts. Figure  2.17(b, d) shows the
degradation of CR with time in the presence of the Ag and Au NPs/GO-G3PAMAM
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 53

Fig. 2.17  UV–Vis kinetics spectra for the degradation of (a) methyl orange and (b) Congo red by
Ag NPs/GO-G3PAMAM and (c-d) by Au NPs/GO-G3PAMAM. Reprinted with permission from
Ref. [123] Copyright @ Royal Society of Chemistry 2014

nanocatalysts, and the catalytic degradation occurs within 180 s and 240 s, respec-
tively. The enhanced degradation observed with the nanocatalysts can be attributed
to the following reasons: (i) it has high surface area of the GO and grafted PAMAM
dendrimer that can adsorb azo dyes, and (ii) NaBH4 is expected to act as the hydride
source and the Ag/Au NP catalysts are expected to activate the azo nitrogen bond
and, also, they could bind with the sulfur and oxygen atoms of the dyes, resulting in
the weakening of the azo double bond via conjugation.

2.5  Summary and Future Prospects

In this book chapter, we summarized the various types of heterojunction photocata-


lysts applicable for environmental remediation, organic dye degradation, wastewa-
ter treatment, air purification, and antibacterial treatments. 1D-0D heterojunctions
made by different types of 1D nanostructures like TiO2 nanotubes, nanobelts, ZnO
nanorods, metal oxide nanowires with 0D metals, and metal oxides of Cu, CuxO,
MnO2, Ag, Ag2O, Au, Pt, etc. have been reported emphasizing the special properties
54 L. R. Nagappagari et al.

associated with the nanostructures like high surface area, abundant active sites, and
efficient charge transfer across the interface of the heterojunction.
Later on, 2D-0D heterojunctions associated with g-C3N4 nanosheets and rGO
and GO nanosheets coupled with metal oxides of Ag2O, Cu2O, ZnO, In2O3, MoO3,
V2O5, and TiO2 and various transition and noble metals of Cu, Ag, Au, Pt, Pd, etc.
were also reported. The important properties associated with the 2D nanosheets like
high surface area and higher electron transport properties and special properties of
nanometals like SPR effect and a greater number of active sites on the surface of the
nanometal particles were highlighted. Various characterization techniques to
investigate the physicochemical properties of those nanomaterials were reported.
For example, HR-TEM, XRD, DRS UV–Vis, PL spectra, time-resolved PL spectra,
transient photocurrent densities, XPS, etc. are studied which were helpful to
investigate the higher photocatalytic efficiencies.
Though the 1D and 2D nanomaterials in association with 0D nanoparticles per-
formed improved efficiencies, they still undergo stability issues when used for
higher time intervals and agglomeration of photocatalysts in the reaction medium.
Thus, it severely limits the practical applications. In order to overcome the stability
issues, much more modifications of the existing materials need to be studied. In
addition, new synthesis techniques should be adopted to make surface protection
layers to avoid the degradation of photocatalysts in the reaction solution when used
for higher timings. Surface modification of photocatalysts needs to be improved so
that a greater number of active sites can be associated with the surface of the
photocatalyst which can readily undergo the redox reactions. Therefore, the
photocatalytic efficiency will be improved and stability issues will be successfully
overcome.

Acknowledgments  Authors acknowledge the National Research Foundation of Korea Grant


funded by the Korean Government (NRF-2019R1l1A3A01041454).

References

1. Shanker, U., Rani, M., & Jassal, V. (2017). Degradation of hazardous organic dyes in water
by nanomaterials. Environmental Chemistry Letters, 15(4), 623–642. https://doi.org/10.1007/
s10311-­017-­0650-­2.
2. Sudrajat, H., & Babel, S. (2016). Rapid photocatalytic degradation of the recalcitrant dye
amaranth by highly active N-WO3. Environmental Chemistry Letters, 14(2), 243–249.
https://doi.org/10.1007/s10311-­015-­0538-­y.
3. Ren, X., Chen, C., Nagatsu, M., & Wang, X. (2011). Carbon nanotubes as adsorbents in
environmental pollution management: a review. Chemical Engineering Journal, 170(2),
395–410. https://doi.org/10.1016/j.cej.2010.08.045.
4. Ahmed, S. N., & Haider, W. (2018). Heterogeneous photocatalysis and its potential applica-
tions in water and wastewater treatment: a review. Nanotechnology, 29(34), 342001. https://
doi.org/10.1088/1361-­6528/aac6ea.
5. Gaya, U. I., & Abdullah, A. H. (2008). Heterogeneous photocatalytic degradation of organic
contaminants over titanium dioxide: a review of fundamentals, progress and problems.
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 55

Journal of Photochemistry and Photobiology C: Photochemistry Review, 9(1), 1–12. https://


doi.org/10.1016/j.jphotochemrev.2007.12.003.
6. Zhao, Z., Dai, Y., Lin, J., & Wang, G. (2014). Highly-ordered mesoporous carbon nitride with
ultrahigh surface area and pore volume as a superior dehydrogenation catalyst. Chemistry of
Materials, 26(10), 3151–3161. https://doi.org/10.1021/cm5005664.
7. Zhao, Y., Hoivik, N., & Wang, K. (2016). Recent advance on engineering titanium dioxide
nanotubes for photochemical and photoelectrochemical water splitting. Nano Energy, 30,
728–744. https://doi.org/10.1016/j.nanoen.2016.09.027.
8. Tian, J., Zhao, Z., Kumar, A., Boughton, R. I., & Liu, H. (2014). Recent progress in design,
synthesis, and applications of one-dimensional TiO2 nanostructured surface heterostruc-
tures: a review. Chemical Society Reviews, 43(20), 6920–6937. https://doi.org/10.1039/
C4CS00180J.
9. Osman, H., Su, Z., & Ma, X. (2017). Efficient photocatalytic degradation of rhodamine B
dye using ZnO/graphitic C3N4 nanocomposites synthesized by microwave. Environmental
Chemistry Letters, 15(3), 435–441. https://doi.org/10.1007/s10311-­017-­0604-­8.
10. Lakshmanareddy, N., Navakoteswara Rao, V., Cheralathan, K.  K., Subramaniam, E.  P., &
Shankar, M.  V. (2019). Pt/TiO2 Nanotube photocatalyst—effect of synthesis methods on
valance state of Pt and its influence on hydrogen production and dye degradation. Journal of
Colloid and Interface Science, 538, 83–98. https://doi.org/10.1016/j.jcis.2018.11.077.
11. Mueller, U., Schubert, M., Teich, F., Puetter, H., Schierle-Arndt, K., & Pastré, J. (2006).
Metal–organic frameworks—prospective industrial applications. Journal of Materials
Chemistry, 16(7), 626–636. https://doi.org/10.1039/B511962F.
12. Taseidifar, M., Makavipour, F., Pashley, R. M., & Rahman, A. F. M. M. (2017). Removal of
heavy metal ions from water using ion flotation. Environmental Technology and Innovation,
8, 182–190. https://doi.org/10.1016/j.eti.2017.07.002.
13. Turhanen, P. A., Vepsäläinen, J. J., & Peräniemi, S. (2015). Advanced material and approach
for metal ions removal from aqueous solutions. Scientific Reports, 5(1), 8992. https://doi.
org/10.1038/srep08992.
14. Chinnakoti, P., Vankayala, R. K., Chunduri, A. L. A., Nagappagari, L. R., Muthukonda, S. V., &
Kamisetti, V. (2016). Trititanate Nanotubes as highly efficient adsorbent for fluoride removal
from water: adsorption performance and uptake mechanism. Journal of Environmental
Chemical Engineering, 4(4), 4754–4768. https://doi.org/10.1016/j.jece.2016.11.007.
15. Chinnakoti, P., Chunduri, A. L. A., Vankayala, R. K., Patnaik, S., & Kamisetti, V. (2017).
Enhanced fluoride adsorption by nano crystalline γ-alumina: adsorption kinetics, isotherm
modeling and thermodynamic studies. Applied Water Science, 7(5), 2413–2423. https://doi.
org/10.1007/s13201-­016-­0437-­9.
16. Singh, J., Singh, P., & Singh, A. (2016). Fluoride ions vs removal technologies: a study.
Arabian Journal of Chemistry, 9(6), 815–824. https://doi.org/10.1016/j.arabjc.2014.06.005.
17. Lang, X., Chen, X., & Zhao, J. (2014). Heterogeneous visible light photocatalysis for
selective organic transformations. Chemical Society Reviews, 43(1), 473–486. https://doi.
org/10.1039/C3CS60188A.
18. Shiraishi, Y., & Hirai, T. (2008). Selective organic transformations on titanium oxide-based
photocatalysts. Journal of Photochemistry and Photobiology C: Photochemistry Review,
9(4), 157–170. https://doi.org/10.1016/j.jphotochemrev.2008.05.001.
19. Wang, C., & Lu, Z. (2015). Catalytic enantioselective organic transformations via visible
light photocatalysis. Organic Chemistry Frontiers, 2(2), 179–190. https://doi.org/10.1039/
C4QO00306C.
20. Wang, W., Niu, Q., Zeng, G., Zhang, C., Huang, D., Shao, B., Zhou, C., Yang, Y., Liu, Y.,
Guo, H., et al. (2020). 1D Porous tubular G-C3N4 capture black phosphorus quantum dots as
1D/0D metal-free photocatalysts for oxytetracycline hydrochloride degradation and hexava-
lent chromium reduction. Applied Catalysis B: Environmental, 273, 119051. https://doi.
org/10.1016/j.apcatb.2020.119051.
56 L. R. Nagappagari et al.

21. Wang, K., Zhang, G., Li, J., Li, Y., & Wu, X. (2017). 0D/2D Z-Scheme heterojunctions of
bismuth tantalate quantum dots/ultrathin g-C3N4 nanosheets for highly efficient visible
light photocatalytic degradation of antibiotics. ACS Applied Materials & Interfaces, 9(50),
43704–43715. https://doi.org/10.1021/acsami.7b14275.
22. Gao, H., Yang, H., Xu, J., Zhang, S., & Li, J. (2018). Strongly coupled G-C3N4 Nanosheets-­
Co3O4 quantum dots as 2D/0D heterostructure composite for peroxymonosulfate activation.
Small, 14(31), 1801353. https://doi.org/10.1002/smll.201801353.
23. Liu, R., Duay, J., & Lee, S. B. (2011). Heterogeneous nanostructured electrode materials for
electrochemical energy storage. Chemical Communications, 47(5), 1384–1404. https://doi.
org/10.1039/C0CC03158E.
24. Lakshmana Reddy, N., Cheralathan, K. K., Durga Kumari, V., Neppolian, B., Muthukonda
Venkatakrishnan, S., Cheralathan, K. K., Durga Kumari, V., Neppolian, B., & Muthukonda
Venkatakrishnan, S. (2018). Photocatalytic reforming of bio-mass derived crude glycerol in
water: a sustainable approach for improved hydrogen generation using Ni(OH)2 decorated
TiO2 nanotubes under solar light irradiation. ACS Sustainable Chemistry & Engineering,
6(3), 3754–3764. https://doi.org/10.1021/acssuschemeng.7b04118.
25. Reddy, N. L., Rao, V. N., Vijayakumar, M., Santhosh, R., Anandan, S., Karthik, M., Shankar,
M. V., Reddy, K. R., Shetti, N. P., Nadagouda, M. N., et al. (2019). A review on Frontiers in
plasmonic nano-photocatalysts for hydrogen production. International Journal of Hydrogen
Energy, 44(21), 10453–10472. https://doi.org/10.1016/j.ijhydene.2019.02.120.
26. Kumar, D.  P., Reddy, N.  L., Karthik, M., Neppolian, B., Madhavan, J., & Shankar,
M. V. V. (2016). Solar light sensitized P-Ag2O/n-TiO2 nanotubes heterojunction photocata-
lysts for enhanced hydrogen production in aqueous-glycerol solution. Solar Energy Materials
& Solar Cells, 154, 78–87. https://doi.org/10.1016/j.solmat.2016.04.033.
27. Praveen Kumar, D., Lakshmana Reddy, N., Karthikeyan, M., Chinnaiah, N., Bramhaiah, V.,
Durga Kumari, V., & Shankar, M. V. (2016). Synergistic effect of nanocavities in anatase TiO2
nanobelts for photocatalytic degradation of methyl orange dye in aqueous solution. Journal
of Colloid and Interface Science, 477, 201–208. https://doi.org/10.1016/j.jcis.2016.05.014.
28. Zhou, M., Wang, S., Yang, P., Luo, Z., Yuan, R., Asiri, A. M., Wakeel, M., & Wang, X. (2018).
Layered heterostructures of ultrathin polymeric carbon nitride and ZnIn2S4 nanosheets for
photocatalytic CO2 reduction. Chemistry  - A European Journal, 24(69), 18529–18534.
https://doi.org/10.1002/chem.201803250.
29. Niu, P., Zhang, L., Liu, G., & Cheng, H.-M. (2012). Graphene-like carbon nitride nanosheets
for improved photocatalytic activities. Advanced Functional Materials, 22(22), 4763–4770.
https://doi.org/10.1002/adfm.201200922.
30. Kale, S. B., Kalubarme, R. S., Mahadadalkar, M. A., Jadhav, H. S., Bhirud, A. P., Ambekar,
J.  D., Park, C.-J., & Kale, B.  B. (2015). Hierarchical 3D ZnIn2S4/graphene nano-­
heterostructures: their in situ fabrication with dual functionality in solar hydrogen produc-
tion and as anodes for lithium ion batteries. Physical Chemistry Chemical Physics, 17(47),
31850–31861. https://doi.org/10.1039/C5CP05546F.
31. Ayekoe, P. Y., Robert, D., & Goné, D. L. (2016). Preparation of effective TiO2/Bi2O3 photo-
catalysts for water treatment. Environmental Chemistry Letters, 14(3), 387–393. https://doi.
org/10.1007/s10311-­016-­0565-­3.
32. Cheng, L., Tian, Y., & Zhang, J. (2018). Construction of P-n heterojunction film of Cu2O/α-­-
Fe2O3 for efficiently photoelectrocatalytic degradation of oxytetracycline. Journal of Colloid
and Interface Science, 526, 470–479. https://doi.org/10.1016/j.jcis.2018.04.106.
33. Fei, W., Li, H., Li, N., Chen, D., Xu, Q., Li, H., He, J., & Lu, J. (2020). Facile fabrica-
tion of ZnO/MoS2 p-n junctions on Ni foam for efficient degradation of organic pollutants
through photoelectrocatalytic process. Solar Energy, 199, 164–172. https://doi.org/10.1016/j.
solener.2020.02.037.
34. Yang, L., Luo, S., Li, Y., Xiao, Y., Kang, Q., & Cai, Q. (2010). High efficient photocata-
lytic degradation of P-nitrophenol on a unique Cu2O/TiO2 p-n heterojunction network cata-
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 57

lyst. Environmental Science & Technology, 44(19), 7641–7646. https://doi.org/10.1021/


es101711k.
35. Huang, H., Liu, C., Ou, H., Ma, T., & Zhang, Y. (2019). Self-sacrifice transformation for
fabrication of Type-I and Type-II heterojunctions in hierarchical BixOyIz/g-C3N4 for effi-
cient visible-light photocatalysis. Applied Surface Science, 470, 1101–1110. https://doi.
org/10.1016/j.apsusc.2018.11.193.
36. Bhoi, Y.  P., & Mishra, B.  G. (2018). Photocatalytic degradation of alachlor using Type-II
CuS/BiFeO3 heterojunctions as novel photocatalyst under visible light irradiation. Chemical
Engineering Journal, 344, 391–401. https://doi.org/10.1016/j.cej.2018.03.094.
37. Shen, J., Shen, J., Zhang, W., Yu, X., Tang, H., Zhang, M., Zulfiqar, & Liu, Q. (2019). Built-in
electric field induced CeO2/Ti3C2-MXene Schottky-junction for coupled photocatalytic
tetracycline degradation and CO2 reduction. Ceramics International, 45(18, Part A),
24146–24153. https://doi.org/10.1016/j.ceramint.2019.08.123.
38. Askari, N., Beheshti, M., Mowla, D., & Farhadian, M. (2020). Synthesis of CuWO4/
Bi2S3 Z-Scheme heterojunction with enhanced cephalexin photodegradation. Journal of
Photochemistry and Photobiology A: Chemistry, 394, 112463. https://doi.org/10.1016/j.
jphotochem.2020.112463.
39. Li, B., Lai, C., Zeng, G., Qin, L., Yi, H., Huang, D., Zhou, C., Liu, X., Cheng, M., Xu, P.,
et al. (2018). Facile hydrothermal synthesis of Z-Scheme Bi2Fe4O9/Bi2WO6 heterojunction
photocatalyst with enhanced visible light photocatalytic activity. ACS Applied Materials &
Interfaces, 10(22), 18824–18836. https://doi.org/10.1021/acsami.8b06128.
40. Murugesan, P., Moses, J. A., & Anandharamakrishnan, C. (2019). Photocatalytic disinfection
efficiency of 2D structure graphitic carbon nitride-based nanocomposites: a review. Journal
of Materials Science, 54(19), 12206–12235. https://doi.org/10.1007/s10853-­019-­03695-­2.
41. Gusain, R., Gupta, K., Joshi, P., & Khatri, O. P. (2019). Adsorptive removal and photocatalytic
degradation of organic pollutants using metal oxides and their composites: a comprehensive
review. Advances in Colloid and Interface Science, 272, 102009. https://doi.org/10.1016/j.
cis.2019.102009.
42. Ma, Y., Wang, X. L., Jia, Y. S., Chen, X. B., Han, H. X., & Li, C. (2014). Titanium dioxide-­
based nanomaterials for photocatalytic fuel generations. Chemical Reviews, 114(19),
9987–10043.
43. Devi, L.  G., & Kavitha, R. (2013). A review on non metal ion doped titania for the
photocatalytic degradation of organic pollutants under UV/solar light: role of photogenerated
charge carrier dynamics in enhancing the activity. Applied Catalysis B: Environmental,
140–141, 559–587. https://doi.org/10.1016/j.apcatb.2013.04.035.
44. Schneider, J., Matsuoka, M., Takeuchi, M., Zhang, J., Horiuchi, Y., Anpo, M., & Bahnemann,
D.  W. (2014). Understanding TiO2 photocatalysis: mechanisms and materials. Chemical
Reviews, 114(19), 9919–9986. https://doi.org/10.1021/cr5001892.
45. Zhao, J., Nan, J., Zhao, Z., Li, N., Liu, J., & Cui, F. (2017). Energy-efficient fabrication of
a novel multivalence Mn3O4-MnO2 heterojunction for dye degradation under visible light
irradiation. Applied Catalysis B: Environmental, 202, 509–517. https://doi.org/10.1016/j.
apcatb.2016.09.065.
46. Nevárez-Martínez, M. C., Kobylański, M. P., Mazierski, P., Wółkiewicz, J., Trykowski, G.,
Malankowska, A., Kozak, M., Espinoza-Montero, P.  J., & Zaleska-Medynska, A. (2017).
Self-organized TiO2-MnO2 nanotube arrays for efficient photocatalytic degradation of tolu-
ene. Molecules, 22(4). https://doi.org/10.3390/molecules22040564.
47. Ganesh, I., Kumar, P. P., Annapoorna, I., Sumliner, J. M., Ramakrishna, M., Hebalkar, N. Y.,
Padmanabham, G., & Sundararajan, G. (2014). Preparation and characterization of Cu-Doped
TiO2 materials for electrochemical, photoelectrochemical, and photocatalytic applications.
Applied Surface Science, 293, 229–247. https://doi.org/10.1016/j.apsusc.2013.12.140.
48. Lee, S. S., Bai, H., Liu, Z., & Sun, D. D. (2013). Novel-structured electrospun TiO2/CuO
composite nanofibers for high efficient photocatalytic cogeneration of clean water and
58 L. R. Nagappagari et al.

energy from dye wastewater. Water Research, 47(12), 4059–4073. https://doi.org/10.1016/j.


watres.2012.12.044.
49. Wang, X., Wu, H.-F., Kuang, Q., Huang, R.-B., Xie, Z.-X., & Zheng, L.-S. (2010). Shape-­
dependent antibacterial activities of Ag2O polyhedral particles. Langmuir, 26(4), 2774–2778.
https://doi.org/10.1021/la9028172.
50. Ren, H.-T., Jia, S.-Y., Wu, Y., Wu, S.-H., Zhang, T.-H., & Han, X. (2014). Improved pho-
tochemical reactivities of Ag2O/g-C3N4  in phenol degradation under UV and Visible
light. Industrial and Engineering Chemistry Research, 53(45), 17645–17653. https://doi.
org/10.1021/ie503312x.
51. Lyu, L.-M., & Huang, M. H. (2011). Investigation of relative stability of different facets of
Ag2O nanocrystals through face-selective etching. Journal of Physical Chemistry C, 115(36),
17768–17773. https://doi.org/10.1021/jp2059479.
52. Sarkar, D., Ghosh, C. K., Mukherjee, S., & Chattopadhyay, K. K. (2013). Three dimensional
Ag2O/TiO2 Type-II (p–n) nanoheterojunctions for superior photocatalytic activity. ACS
Applied Materials & Interfaces, 5(2), 331–337. https://doi.org/10.1021/am302136y.
53. Morales, M. V., Asedegbega-Nieto, E., Iglesias-Juez, A., Rodríguez-Ramos, I., & Guerrero-­
Ruiz, A. (2015). Role of exposed surfaces on zinc oxide nanostructures in the catalytic ethanol
transformation. ChemSusChem, 8(13), 2223–2230. https://doi.org/10.1002/cssc.201500425.
54. Wang, Y., Shi, R., Lin, J., & Zhu, Y. (2011). Enhancement of photocurrent and photocatalytic
activity of ZnO hybridized with graphite-like C3N4. Energy & Environmental Science, 4(8),
2922–2929. https://doi.org/10.1039/C0EE00825G.
55. Zha, R., Nadimicherla, R., & Guo, X. (2015). Ultraviolet photocatalytic degradation of
methyl orange by nanostructured TiO2/ZnO heterojunctions. Journal of Materials Chemistry
A, 3(12), 6565–6574. https://doi.org/10.1039/C5TA00764J.
56. Zheng, X., Li, D., Li, X., Chen, J., Cao, C., Fang, J., Wang, J., He, Y., & Zheng, Y. (2015).
Construction of ZnO/TiO2 photonic crystal heterostructures for enhanced photocatalytic
properties. Applied Catalysis B: Environmental, 168–169, 408–415. https://doi.org/10.1016/j.
apcatb.2015.01.001.
57. Sun, W., Meng, S., Zhang, S., Zheng, X., Ye, X., Fu, X., & Chen, S. (2018). Insight into the
transfer mechanisms of photogenerated carriers for heterojunction photocatalysts with the
analogous positions of valence band and conduction band: a case study of ZnO/TiO2. Journal
of Physical Chemistry C, 122(27), 15409–15420. https://doi.org/10.1021/acs.jpcc.8b03753.
58. Deng, H., He, H., Sun, S., Zhu, X., Zhou, D., Han, F., Huang, B., & Pan, X. (2019).
Photocatalytic degradation of dye by Ag/TiO2 nanoparticles prepared with different sol–
gel crystallization in the presence of effluent organic matter. Environmental Science and
Pollution Research, 26(35), 35900–35912. https://doi.org/10.1007/s11356-­019-­06728-­0.
59. Scuderi, V., Impellizzeri, G., Romano, L., Scuderi, M., Nicotra, G., Bergum, K.,
Irrera, A., Svensson, B.  G., & Privitera, V. (2014). TiO2-Coated nanostructures for
dye photo-­ degradation in water. Nanoscale Research Letters, 9(1), 458. https://doi.
org/10.1186/1556-276X-9-458.
60. Fei, J., & Li, J. (2015). Controlled preparation of porous TiO2–Ag nanostructures through
supramolecular assembly for plasmon-enhanced photocatalysis. Advanced Materials, 27(2),
314–319. https://doi.org/10.1002/adma.201404007.
61. Wen, Y., Liu, B., Zeng, W., & Wang, Y. (2013). Plasmonic photocatalysis properties of
Au nanoparticles precipitated anatase/rutile mixed TiO2 nanotubes. Nanoscale, 5(20),
9739–9746. https://doi.org/10.1039/C3NR03024E.
62. Zhang, Z., Baek, M., Song, H., & Yong, K. (2017). An unconventional outer-to-inner synthesis
strategy for core (Au)-Shell nanostructures with photo-electrochemical enhancement.
Nanoscale, 9(16), 5342–5351. https://doi.org/10.1039/C7NR00336F.
63. Bian, Z., Tachikawa, T., Zhang, P., Fujitsuka, M., & Majima, T. A. (2014). TiO2 Superstructure-­
based plasmonic photocatalysts exhibiting efficient charge separation and unprecedented
activity. Journal of the American Chemical Society, 136(1), 458–465. https://doi.org/10.1021/
ja410994f.
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 59

64. Yu, Y., Wen, W., Qian, X.-Y., Liu, J.-B., & Wu, J.-M. (2017). UV and Visible light photocata-
lytic activity of Au/TiO2 nanoforests with anatase/rutile phase junctions and controlled Au
locations. Scientific Reports, 7(1), 41253. https://doi.org/10.1038/srep41253.
65. Do, T. C. M. V., Nguyen, D. Q., Nguyen, K. T., & Le, P. H. (2019). TiO(2) and Au-TiO(2)
Nanomaterials for rapid photocatalytic degradation of antibiotic residues in aquaculture
wastewater. Mater (Basel, Switzerland), 12(15), 2434. https://doi.org/10.3390/ma12152434.
66. Yang, Z., Lu, J., Ye, W., Yu, C., & Chang, Y. (2017). Preparation of Pt/TiO2 hollow nanofi-
bers with highly visible light photocatalytic activity. Applied Surface Science, 392, 472–480.
https://doi.org/10.1016/j.apsusc.2016.09.065.
67. Cha, G., Schmuki, P., & Altomare, M. (2017). Anodic TiO2 nanotube membranes: site-­
selective Pt-activation and photocatalytic H2 evolution. Electrochimica Acta, 258, 302–310.
https://doi.org/10.1016/j.electacta.2017.11.030.
68. Fornari, A., Maria, D., de Araujo, M.  B., Bergamin, D.  C., Machado, G., Teixeira, S.  R.,
& Weibel, D. E. (2016). Photocatalytic reforming of aqueous formaldehyde with hydrogen
generation over TiO2 nanotubes loaded with Pt or Au nanoparticles. International Journal of
Hydrogen Energy, 41(27), 11599–11607. https://doi.org/10.1016/j.ijhydene.2016.02.055.
69. Slamet; Nasution, H. W., Purnama, E., Kosela, S., Gunlazuardi, J. Photocatalytic reduction of
CO2 on Copper-Doped titania catalysts prepared by improved-impregnation method. Catalysis
Communications 2005, 6 (5), 313–319. doi: https://doi.org/10.1016/j.catcom.2005.01.011.
70. Rao, V. N., Reddy, N. L., Kumari, M. M., Cheralathan, K. K., Ravi, P., Sathish, M., Neppolian,
B., Reddy, K. R., Shetti, N. P., Prathap, P., et al. (2019). Sustainable hydrogen production for
the Greener environment by quantum dots-based efficient photocatalysts: a review. Journal
of Environmental Management. Academic Press, October 15. https://doi.org/10.1016/j.
jenvman.2019.07.017.
71. Onsuratoom, S., Puangpetch, T., & Chavadej, S. (2011). Comparative investigation of
hydrogen production over Ag-, Ni-, and Cu-loaded mesoporous-assembled TiO2-ZrO2 mixed
oxide nanocrystal photocatalysts. Chemical Engineering Journal, 173(2), 667–675. https://
doi.org/10.1016/j.cej.2011.08.016.
72. Bessekhouad, Y., Robert, D., & Weber, J. V. (2005). Photocatalytic activity of Cu2O/TiO2,
Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions. Catalysis Today, 101(3–4), 315–321. https://
doi.org/10.1016/j.canod.2005.03.038.
73. Sang, L., Zhang, S., Zhang, J., Yu, Z., Bai, G., & Du, C. (2019). TiO2 Nanotube arrays deco-
rated with plasmonic Cu, CuO nanoparticles, and Eosin Y dye as efficient photoanode for
water splitting. Materials Chemistry and Physics, 231, 27–32. https://doi.org/10.1016/j.
matchemphys.2019.04.018.
74. Liang, M., Lei, Q., Sun, S., & Zhu, Y. (2020). Plasmonic responses of Cu–Ag bimetallic
system: influence of distinctiveness and arrangements. Optical Materials (Amst), 100,
109655. https://doi.org/10.1016/j.optmat.2020.109655.
75. Lv, X.-J., Zhou, S.-X., Zhang, C., Chang, H.-X., Chen, Y., & Fu, W.-F. (2012). Synergetic
effect of Cu and graphene as cocatalyst on TiO2 for enhanced photocatalytic hydrogen evolu-
tion from solar water splitting. Journal of Materials Chemistry, 22(35), 18542–18549. https://
doi.org/10.1039/C2JM33325B.
76. Meriam Suhaimy, S. H., Abd Hamid, S. B., Lai, C. W., Hasan, M. R., & Johan, M. R. (2016).
TiO2 Nanotubes supported Cu nanoparticles for improving photocatalytic degradation of
simazine under UV illumination. Catalysts, 6(11), 167. https://doi.org/10.3390/catal6110167.
77. Kumar, S., Reddy, N. L., Kumar, A., Shankar, M. V., Krishnan, V., Kumar, S., Lakshmana,
N., Reddy, A. K., & Muthukonda Venkatakrishnan Shankar, V. K. (2017). Two dimensional
N-Doped ZnO-graphitic carbon nitride nanosheets heterojunctions with enhanced photo-
catalytic hydrogen evolution. International Journal of Hydrogen Energy, 43(8), 3988–4002.
https://doi.org/10.1016/j.ijhydene.2017.09.113.
78. Zeng, D., Xu, W., Ong, W.-J., Xu, J., Ren, H., Chen, Y., Zheng, H., & Peng, D.-L. (2018).
Toward Noble-metal-free visible-light-driven photocatalytic hydrogen evolution: monodis-
perse sub–15  nm Ni2P nanoparticles anchored on porous g-C3N4 nanosheets to engineer
60 L. R. Nagappagari et al.

0D-2D heterojunction interfaces. Applied Catalysis B: Environmental, 221, 47–55. https://


doi.org/10.1016/j.apcatb.2017.08.041.
79. Abdelhafeez, I. A., Yao, Q., Wang, C., Su, Y., Zhou, X., & Zhang, Y. (2019). Green synthesis
of ultrathin edge-activated foam-like carbon nitride nanosheets for enhanced photocatalytic
performance under visible light irradiation. Sustain. Energy Fuels, 3(7), 1764–1775. https://
doi.org/10.1039/C9SE00263D.
80. Kumar Suneel, A., Reddy, N.  L., Kushwaha, H.  S., & Kumar. (2017). Efficient electron
transfer across ZnO-MoS 2 -RGO heterojunction for remarkably enhanced sunlight
driven photocatalytic hydrogen evolution. ChemSusChem, 10(18), 3588–3603. https://doi.
org/10.1002/cssc.201701024.
81. Luo, B., Liu, G., & Wang, L. (2016). Recent advances in 2D materials for photocatalysis.
Nanoscale, 8(13), 6904–6920. https://doi.org/10.1039/C6NR00546B.
82. Ong, W.-J., Tan, L.-L., Ng, Y. H., Yong, S.-T., & Chai, S.-P. (2016). Graphitic carbon nitride
(g-C3N4)-based photocatalysts for artificial photosynthesis and environmental remediation:
are we a step closer to achieving sustainability? Chemical Reviews, 116(12), 7159–7329.
https://doi.org/10.1021/acs.chemrev.6b00075.
83. Perreault, F., de Faria, A., & Elimelech, M. (2015). Environmental applications of graphene-­
based nanomaterials. Chemical Society Reviews, 44(16), 5861–5896. https://doi.org/10.1039/
C5CS00021A.
84. Li, F., Jiang, X., Zhao, J., & Zhang, S. (2015). Graphene oxide: a promising nanomaterial for
energy and environmental applications. Nano Energy, 16, 488–515. https://doi.org/10.1016/j.
nanoen.2015.07.014.
85. Wen, J., Li, X., Li, H., Ma, S., He, K., Xu, Y., Fang, Y., Liu, W., & Gao, Q. (2015). Enhanced
visible-light H2 evolution of g-C3N4 photocatalysts via the synergetic effect of amorphous
NiS and cheap metal-free carbon black nanoparticles as co-catalysts. Applied Surface
Science, 358, 204–212. https://doi.org/10.1016/j.apsusc.2015.08.244.
86. Stan, M. S., Badea, M. A., Pircalabioru, G. G., Chifiriuc, M. C., Diamandescu, L., Dumitrescu,
I., Trica, B., Lambert, C., & Dinischiotu, A. (2019). Designing cotton fibers impregnated
with photocatalytic graphene oxide/Fe, N-Doped TiO2 particles as prospective industrial self-­
cleaning and biocompatible textiles. Materials Science and Engineering: C, 94, 318–332.
https://doi.org/10.1016/j.msec.2018.09.046.
87. Kumar, S., Reddy, N. L., Kushwaha, H. S., Kumar, A., Shankar, M. V., Bhattacharyya, K.,
Halder, A., & Krishnan, V. (2017). Efficient electron transfer across a ZnO–MoS2–reduced
graphene oxide heterojunction for enhanced sunlight-driven photocatalytic hydrogen evolu-
tion. ChemSusChem, 10(18), 3588–3603. https://doi.org/10.1002/cssc.201701024.
88. Cao, S.-W., Liu, X.-F., Yuan, Y.-P., Zhang, Z.-Y., Liao, Y.-S., Fang, J., Loo, S. C. J., Sum,
T.  C., & Xue, C. (2014). Solar-to-fuels conversion over In2O3/g-C3N4 hybrid photo-
catalysts. Applied Catalysis B: Environmental, 147, 940–946. https://doi.org/10.1016/j.
apcatb.2013.10.029.
89. Jiang, Z., Jiang, D., Yan, Z., Liu, D., Qian, K., & Xie, J. A. (2015). New visible light active
multifunctional ternary composite based on TiO2–In2O3 nanocrystals heterojunction decorated
porous graphitic carbon nitride for photocatalytic treatment of hazardous pollutant and H2
evolution. Applied Catalysis B: Environmental, 170–171, 195–205. https://doi.org/10.1016/j.
apcatb.2015.01.041.
90. Zang, Y., Li, L., Li, X., Lin, R., & Li, G. (2014). Synergistic collaboration of G-C3N4/SnO2
composites for enhanced visible-light photocatalytic activity. Chemical Engineering Journal,
246, 277–286. https://doi.org/10.1016/j.cej.2014.02.068.
91. Tian, N., Huang, H., He, Y., Guo, Y., Zhang, T., & Zhang, Y. (2015). Mediator-free direct
Z-scheme photocatalytic system: BiVO4/g-C3N4 organic–inorganic hybrid photocatalyst
with highly efficient visible-light-induced photocatalytic activity. Dalton Transactions,
44(9), 4297–4307. https://doi.org/10.1039/C4DT03905J.
92. Li, Y., Wang, J., Yang, Y., Zhang, Y., He, D., An, Q., & Cao, G. (2015). Seed-induced growing
various TiO2 nanostructures on g-C3N4 nanosheets with much enhanced photocatalytic activ-
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 61

ity under visible light. Journal of Hazardous Materials, 292, 79–89. https://doi.org/10.1016/j.
jhazmat.2015.03.006.
93. Chen, Y.-C., Pu, Y.-C., & Hsu, Y.-J. (2012). Interfacial charge carrier dynamics of the three-­
component In2O3–TiO2–Pt heterojunction system. Journal of Physical Chemistry C, 116(4),
2967–2975. https://doi.org/10.1021/jp210033y.
94. Wang, X., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J. M., Domen, K., &
Antonietti, M. A. (2009). Metal-free polymeric photocatalyst for hydrogen production from
water under visible light. Nature Materials, 8(1), 76–80. https://doi.org/10.1038/nmat2317.
95. Kumar, R., Ansari, M.  O., Barakat, M.  A., & Rashid, J. (2019). Chapter 13: Graphene/
metal oxide–based nanocomposite as photocatalyst for degradation of water pollutants. In
M.  Jawaid, A.  Ahmad, & D.  Lokhat (Eds.), Graphene-based nanotechnologies for energy
and environmental applications, micro and nano technologies (pp.  221–240). Elsevier.
https://doi.org/10.1016/B978-­0-­12-­815811-­1.00013-­2.
96. Rakkesh, R.  A., Durgalakshmi, D., & Balakumar, S. (2018). Beyond chemical bonding
interaction: an insight into the growth process of 1D ZnO on few-layer graphene for excel-
lent photocatalytic and room temperature gas sensing applications. ChemistrySelect, 3(25),
7302–7309. https://doi.org/10.1002/slct.201800987.
97. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2015). Nanostructuring of a GNS-­
V2O5–TiO2 Core–Shell photocatalyst for water remediation applications under sun-light irra-
diation. RSC Advances, 5(24), 18633–18641. https://doi.org/10.1039/C5RA00180C.
98. Li, B., & Cao, H. (2011). ZnO@graphene composite with enhanced performance for the
removal of dye from water. Journal of Materials Chemistry, 21(10), 3346–3349. https://doi.
org/10.1039/C0JM03253K.
99. Wang, Y., Wang, W., Mao, H., Lu, Y., Lu, J., Huang, J., Ye, Z., & Lu, B. (2014). Electrostatic
self-assembly of BiVO4–reduced graphene oxide nanocomposites for highly efficient vis-
ible light photocatalytic activities. ACS Applied Materials & Interfaces, 6(15), 12698–12706.
https://doi.org/10.1021/am502700p.
100. Ajay Rakkesh, R., Durgalakshmi, D., Karthe, P., & Balakumar, S. (2019). Role of interfacial
charge transfer process in the graphene-ZnO-MoO3 Core-Shell nanoassemblies for efficient
disinfection of industrial effluents. Processing and Application of Ceramics, 13(4), 376–386.
https://doi.org/10.2298/PAC1904376A.
101. Kumar, S., Sharma, R., Sharma, V., Harith, G., Sivakumar, V., & Krishnan, V. (2016). Role
of RGO support and irradiation source on the photocatalytic activity of CdS–ZnO semi-
conductor nanostructures. Beilstein Journal of Nanotechnology, 7, 1684–1697. https://doi.
org/10.3762/bjnano.7.161.
102. Gurushantha, K., Anantharaju, K.  S., Renuka, L., Sharma, S.  C., Nagaswarupa, H.  P.,
Prashantha, S. C., Vidya, Y. S., & Nagabhushana, H. (2017). New Green synthesized reduced
graphene oxide–ZrO2 composite as high performance photocatalyst under sunlight. RSC
Advances, 7(21), 12690–12703. https://doi.org/10.1039/C6RA25823A.
103. Jiao, T., Zhao, H., Zhou, J., Zhang, Q., Luo, X., Hu, J., Peng, Q., & Yan, X. (2015). Self-­
Assembly reduced graphene oxide nanosheet hydrogel fabrication by anchorage of chitosan/
silver and its potential efficient application toward dye degradation for wastewater treatments.
ACS Sustainable Chemistry & Engineering, 3(12), 3130–3139. https://doi.org/10.1021/
acssuschemeng.5b00695.
104. An, S. J., Zhu, Y., Lee, S. H., Stoller, M. D., Emilsson, T., Park, S., Velamakanni, A., An, J.,
& Ruoff, R. S. (2010). Thin film fabrication and simultaneous anodic reduction of deposited
graphene oxide platelets by electrophoretic deposition. Journal of Physical Chemistry
Letters, 1(8), 1259–1263. https://doi.org/10.1021/jz100080c.
105. Wang, H., Robinson, J. T., Li, X., & Dai, H. (2009). Solvothermal reduction of chemically
exfoliated graphene sheets. Journal of the American Chemical Society, 131(29), 9910–9911.
https://doi.org/10.1021/ja904251p.
106. Mady, A.  H., Baynosa, M.  L., Tuma, D., & Shim, J.-J. (2017). Facile microwave-assisted
Green synthesis of Ag-ZnFe2O4@rGO nanocomposites for efficient removal of organic dyes
62 L. R. Nagappagari et al.

under UV- and Visible-light irradiation. Applied Catalysis B: Environmental, 203, 416–427.
https://doi.org/10.1016/j.apcatb.2016.10.033.
107. Gupta, A., Jamatia, R., Patil, R. A., Ma, Y.-R., & Pal, A. K. (2018). Copper Oxide/reduced
graphene oxide nanocomposite-catalyzed synthesis of flavanones and flavanones with tri-
azole hybrid molecules in one pot: a green and sustainable approach. ACS Omega, 3(7),
7288–7299. https://doi.org/10.1021/acsomega.8b00334.
108. Sree, G.  S., Botsa, S.  M., Reddy, B.  J. M., & Ranjitha, K.  V. B. (2020). Enhanced UV–
Visible triggered photocatalytic degradation of brilliant green by reduced graphene oxide
based NiO and CuO ternary nanocomposite and their antimicrobial activity. Arabian Journal
of Chemistry, 13(4), 5137–5150. https://doi.org/10.1016/j.arabjc.2020.02.012.
109. Mandal, S.  K., Dutta, K., Pal, S., Mandal, S., Naskar, A., Pal, P.  K., Bhattacharya, T.  S.,
Singha, A., Saikh, R., De, S., et al. (2019). Engineering of ZnO/RGO nanocomposite pho-
tocatalyst towards rapid degradation of toxic dyes. Materials Chemistry and Physics, 223,
456–465. https://doi.org/10.1016/j.matchemphys.2018.11.002.
110. Lakshmana Reddy, N., Navakoteswara Rao, V., Mamatha Kumari, M., Kakarla,
R. R. R. R., Ravi, P., Sathish, M., Karthik, M., Venkatakrishnan, M., & Inamuddin, S. (2018).
Nanostructured semiconducting materials for efficient hydrogen generation. Environmental
Chemistry Letters, 16(3), 765–796. https://doi.org/10.1007/s10311-­018-­0722-­y.
111. Jose, D., Sorensen, C. M., Rayalu, S. S., Shrestha, K. M., & Klabunde, K. J. (2013). Au-TiO2
Nanocomposites and efficient photocatalytic hydrogen production under UV–Visible and
visible light illuminations: a comparison of different crystalline forms of TiO2. International
Journal of Photoenergy, 685614(10). https://doi.org/10.1155/2013/685614.
112. Phivilay, S. P., Puretzky, A. A., Domen, K., & Wachs, I. E. (2013). Nature of catalytic active
sites present on the surface of advanced bulk tantalum mixed oxide photocatalysts. ACS
Catalysis, 3(12), 2920–2929. https://doi.org/10.1021/cs400662m.
113. Zhang, H., Wei, J., Dong, J., Liu, G., Shi, L., An, P., Zhao, G., Kong, J., Wang, X., Meng,
X., et  al. (2016). Efficient visible-light-driven carbon dioxide reduction by a single-atom
implanted metal–organic framework. Angewandte Chemie International Edition, 55(46),
14310–14314. https://doi.org/10.1002/anie.201608597.
114. Khan, I., Saeed, K., & Khan, I. (2019). Nanoparticles properties, applications and toxicities.
Arabian Journal of Chemistry, 12(7), 908–931. https://doi.org/10.1016/j.arabjc.2017.05.011.
115. Faisal, M., Jalalah, M., Harraz, F. A., El-Toni, A. M., Khan, A., & Al-Assiri, M. S. (2020).
Au nanoparticles-doped g-C3N4 nanocomposites for enhanced photocatalytic perfor-
mance under visible light illumination. Ceramics International. https://doi.org/10.1016/j.
ceramint.2020.05.250.
116. Wang, H., Wang, G., Zhang, Y., Ma, Y., Wu, Z., Gao, D., Yang, R., Wang, B., Qi, X., &
Yang, J. (2019). Preparation of RGO/TiO2/Ag aerogel and its photodegradation perfor-
mance in gas phase formaldehyde. Scientific Reports, 9(1), 16314. https://doi.org/10.1038/
s41598-­019-­52541-­7.
117. Huang, L., Yang, H., Zhang, Y., & Xiao, W. (2016). Study on synthesis and antibacterial prop-
erties of Ag NPs/GO nanocomposites. Journal of Nanomaterials, 2016, 5685967. https://doi.
org/10.1155/2016/5685967.
118. Li, L., Chen, M., Huang, G., Yang, N., Zhang, L., Wang, H., Liu, Y., Wang, W., & Gao,
J. (2014). A Green method to prepare Pd–Ag nanoparticles supported on reduced graphene
oxide and their electrochemical catalysis of methanol and ethanol oxidation. Journal of
Power Sources, 263, 13–21. https://doi.org/10.1016/j.jpowsour.2014.04.021.
119. Shi, L., Li, Z., Marcus, K., Wang, G., Liang, K., Niu, W., & Yang, Y. (2018). Integration of Au
nanoparticles with a G-C3N4 based heterostructure: switching charge transfer from Type-II
to Z-scheme for enhanced visible light photocatalysis. Chemical Communications, 54(30),
3747–3750. https://doi.org/10.1039/C8CC01370E.
120. Kasturi, S., Torati, S.  R., Eom, Y.  J., Ahmad, S., Lee, B.-J., Yu, J.-S., & Kim, C. (2020).
Real-time monitored photocatalytic activity and electrochemical performance of an RGO/
2  Nanostructured Heterojunction (1D-0D and 2D-0D) Photocatalysts… 63

Pt nanocomposite synthesized via a Green approach. RSC Advances, 10(23), 13722–13731.


https://doi.org/10.1039/D0RA00541J.
121. Liu, C.-C., Xu, H., Wang, L., & Qin, X. (2017). Facile one-pot green synthesis and antibacte-
rial activities of GO/Ag nanocomposites. Acta Metallurgica Sinica (English Letters), 30(1),
36–44. https://doi.org/10.1007/s40195-­016-­0517-­8.
122. Liu, Y., Tian, C., Yan, B., Lu, Q., Xie, Y., Chen, J., Gupta, R., Xu, Z., Kuznicki, S. M., Liu, Q.,
et al. (2015). Nanocomposites of graphene oxide{,} Ag nanoparticles{,} and magnetic ferrite
nanoparticles for elemental mercury (Hg0) removal. RSC Advances, 5(20), 15634–15640.
https://doi.org/10.1039/C4RA16016A.
123. Rajesh, R., Kumar, S. S., & Venkatesan, R. (2014). Efficient degradation of azo dyes using Ag
and Au nanoparticles stabilized on graphene oxide functionalized with PAMAM dendrimers.
New Journal of Chemistry, 38(4), 1551–1558. https://doi.org/10.1039/C3NJ01050C.
Chapter 3
Hierarchical Nanostructures
for Photocatalytic Applications

R. Ajay Rakkesh, Durgalakshmi Dhinasekaran, M. V. Shankar,


and S. Balakumar

3.1  B
 asic Concepts of Hierarchical Nanostructures
in Photocatalytic Field

Recently, nanosize- and quantum-size-based photocatalysts have attracted immense


attention among the material scientist due to their excellent physicochemical prop-
erties in solving energy- and environmental-related problems [1–3]. The interaction
between light energy and metal oxide semiconductors has generated excitons, and
reactive oxygen species provided a sustainable opportunity to decompose any
organic matter at micro level. Currently, a broad variety of metal oxide semicon-
ducting materials have been demonstrated for specific as well as multifunctional
applications, such as metal oxides with d0 and d10 configurations, semiconductor-­
based plasmonic nanostructures, metal oxynitrides/sulphides, metal-organic frame-
works and perovskite-based photocatalysts [4–8].

R. A. Rakkesh
National Centre for Nanoscience and Nanotechnology, University of Madras,
Chennai, Tamil Nadu, India
Department of Physics and Nanotechnology, SRM Institute of Science and Technology,
Kattankulathur, Tamil Nadu, India
D. Dhinasekaran
Department of Medical Physics, Anna University, Chennai, Tamil Nadu, India
M. V. Shankar
Department of Materials Science and Nanotechnology, Yogi Vemana University,
Kadapa, Andhra Pradesh, India
S. Balakumar (*)
National Centre for Nanoscience and Nanotechnology, University of Madras,
Chennai, Tamil Nadu, India

© Springer Nature Switzerland AG 2021 65


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_3
66 R. A. Rakkesh et al.

A perfect photocatalyst shall satisfy both bulk and surface properties, such as
smaller bandgap, which can proficiently absorb a broad region of the natural sunlight
spectrum, appropriate energy band potential, excellent in the separation of electron–
hole pairs and transport, improved surface-interface with reactants, economically
viable, better stability in a different environment [9, 10]. However, few photocata-
lysts fulfilled some of the above-mentioned criteria to become a better photocatalytic
material for practical applications. With reference to the precious reports, the synthe-
sis processes for the photocatalysts and optimized parameters can significantly
enhance the material properties and photocatalytic activity. As a result, vigorous
research is underway to develop versatile and stable visible light active photocatalyst
with existing synthesis process either by physical integration or chemical bonding
with carbon nanostructures as a co-catalyst or forming interface structure with suit-
able materials and fabricating Z-schemes and so on. Normally, these modifications
in the material aspects have to enhance the photocatalytic performances in two dif-
ferent ways: one is a structural modification and another one is compositional varia-
tions [11, 12]. The structural modification of the photocatalysts is strongly dependent
on the material morphologies as well as the alteration in crystallite/grain structure
features. The compositional variation is to modify the stoichiometric ratios of the
host material by either doping or substitution with metals, nitrides, rare earth ele-
ments, etc., to engineer the bandgap energy [13, 14]. Till date, diverse approaches
were reported to fabricate the required crystallographic features and morphologies
for excellent photocatalytic performance under the visible spectrum of solar light.
Among them, the hierarchical nanostructures with metal oxide semiconductors have
become an appropriate nanostructure to meet the desired criteria due to their unique
morphology, higher surface area with exposed facets and structure-dependent photo-
catalytic performances under visible light irradiation [15, 16] (Fig. 3.1).

Fig. 3.1 Schematic
presentation of hierarchical
photocatalysts and its
applications
3  Hierarchical Nanostructures for Photocatalytic Applications 67

Understanding the photocatalytic performance in terms of designing hierarchical


nanostructure with tuneable morphology, crystalline nature with exposed facets and
desired crystal growth orientations makes a perfect hierarchical structure for effi-
cient activities [17, 18]. Hierarchical nanostructures are generally made up of three-­
dimensional self-assembly of prime nanostructure in the nanometre regime
(spherical nanoparticles, nanorods, porous nanoparticles, nanosheets or nanotubes).
Hierarchical photocatalysts are generally defined as semiconducting nanostructured
materials with different dimensional domains like branched/pores/projection from
the surfaces at multiple levels [19, 20]. Hierarchical nanostructure with numerous
characteristic has been demonstrated recently, such as flowerlike structure, porous
network, branched rods, uracil-like structure and self-organized 3D networks, mim-
icking the bio-inspired structures existing naturally like grass, leaves and trees [21–
23]. Essentially, hierarchical structure provides a space for easily available network
structures for better adsorption property, excellent surface area, high electron trans-
port property due to the projected crystal facets and strong binding support for co-­
catalyst materials. These exceptional characteristics of semiconductor-based
hierarchical nanostructures show futuristic structures for excellent photocatalytic
material [24, 25]. Currently, several publications emphasized the importance of syn-
thesis process in preparation of hierarchical nanostructures. Especially, the photo-
catalytic performance of the hierarchical nanostructures provides excellent
efficiency due to the synergetic effect combined with the multidirectional networks
self-assembled in the nanostructures [26–28].
This chapter portrays the importance of hierarchical nanostructures, fabrication
strategies and significant applications related to energy and environmental fields.
Particularly this chapter summarizes the novel synthesis techniques such as precipi-
tation synthesis, hydrothermal method, solvothermal method, microwave treatment
and metal-organic framework-directed synthesis approach for the preparation of
metal oxide-based hierarchical nanostructures. Finally, some key applications like
photocatalytic water remediation, photocatalytic hydrogen fuel production and pho-
tocatalytic carbon dioxide are summarized.

3.2  Preparation Strategies

The above discussion shows that the vital materialistic characteristics like capability
to take up wide range of solar spectrum and higher surface area of hierarchical
nanostructures play key role in photocatalytic performances. To date, several reports
exist on design and development of hierarchical semiconductor photocatalysts.
Generally, the fabrication strategies are based on physical (lithography, laser-­
induced depositions and vapour deposition), chemical (liquid-phase synthesis
routes) and biological techniques. With respect to the other synthesis routes, liquid-­
phase synthesis routes have great advantages over other routes in terms of cost,
energy consumption, stability and scalable productivity. Moreover, the liquid-phase
synthesis can produce many interesting building blocks in the nano regime
68 R. A. Rakkesh et al.

(including 1D, 2D and 3D networked branches, heterostructures and porous hierar-


chical nanostructures).
The following methods are reported widely for the preparation of metal oxide-­
based hierarchical nanostructures:
• Precipitation method.
• Hydrothermal method.
• Solvothermal method.
• Microwave treatment.
• Metal-organic framework-directed synthesis approach.

3.2.1  Precipitation Method

The precipitation technique is a widely used synthesis process for the fabrication of
semiconductor-based hierarchical nanostructured materials due to its vital merits
over the reaction time, repeatability, stability of the structure, mass production,
crystalline nature, controllable particle size and purity and requires basic facilities.
In a typical synthesis process, various kinds of reducing agent like oxalic acid,
sodium hydroxide, potassium hydroxide, ammonia, urea, etc. and precursor materi-
als have been used. Addition of reducing agents into the metallic precursor dis-
solved in water results in precipitation of the metal. The final precipitate is collected
by the process of centrifugation and filtration. The shape, size and surface architec-
ture of the material can be regulated and modified by altering the reaction process
like the stoichiometric ratio of the reacted products, reaction period, temperature
and solution pH [29–32].
Guzman et  al. reported that snowflake and flowerlike ZnO hierarchical nano-
structures were fabricated by a facile liquid-phase precipitation technique [33].
These ZnO nanoarrays were synthesized by using ZnNO3 as a precursor and sodium
hydroxide as a reducing agent. The effect of temperature has been studied to obtain
two different morphologies at 60  °C and 70  °C for the snowflake and flowerlike
hierarchical nanostructures, respectively. The authors have also evaluated that self-­
aggregation is the only process for morphological variation under the influence of
reaction temperature (Fig. 3.2).
Li et al. demonstrated that nanoflake-array-flower ZnO hierarchical nanostruc-
ture has been synthesized by a novel precipitation technique [34]. The author used
sodium lignosulfonate as a surfactant and was able to create the pores in the flower
region of zinc oxide nanoflake arrays. Based on the experimental findings, the
author reported that formation mechanism of ZnO hierarchical structure attributes
the synergistic effect. The simultaneous self-aggregation and electrostatic interac-
tion of sodium lignosulfonate and ZnO crystals help to form the flowerlike array.
This hierarchical structure facilitates the absorption property of zinc oxide and
improves the layer-by-layer profile of sodium lignosulfonate. Thus, it generates the
flowerlike ZnO to form a self-assembled flaky matrix.
3  Hierarchical Nanostructures for Photocatalytic Applications 69

Fig. 3.2  Pictorial representation of precipitation method of synthesizing hierarchical nanostruc-


ture [76]

3.2.2  Hydrothermal Method

Hydrothermal method is a low-temperature process that facilitates the preparation


of nanostructures and easy to scale up for bulk preparation. To carry out the hydro-
thermal method, Teflon-lined stainless steel autoclave is used under the controlled
pressure and temperature condition in the elevated atmosphere; normally the tem-
perature exceeds in the range between 100 °C to 200 °C. The pressure generated in
the autoclave can be varied by changing the reaction temperature. Thus, it helps to
form the crystalline hierarchical nanostructure at relatively low temperature. Some
of the factors like a concentration of the precursor, temperature, reaction period and
the pH of the chemical reaction can influence the growth and nucleation of the nano-
structures, and consequently, it is promising to influence the shape and size of the
nanomaterial by altering the above-listed factors. Morphological changing agents or
surfactants are also used to modify the morphology of the synthesized nanostruc-
tures. The morphological evolution process can range from spherical nanoparticles
to multidimensional uracil structures upon varying the above-mentioned factors.
The hydrothermal method has numerous advantageous over other techniques like an
eco-friendly way of synthesis, easy removal of by-products, high crystalline nature
and hierarchical morphology [35–37].
Zhu et al. demonstrated the hydrothermal synthesis route for the fabrication of
flowerlike morphology of ZnO hierarchical nanostructure [38]. Moreover, the
author mentioned the significance of using cetyl trimethyl ammonium bromide
(CTAB) as a surfactant in the hydrothermal process, as it helps to acquire proper
70 R. A. Rakkesh et al.

nucleation, growth and self-assembly of zinc oxide crystal for the arrangement of
flowerlike morphology. Zhang et al. developed the rapid hydrothermal route for the
fabrication of ultrathin birnessite-type MnO2 hierarchical nanostructures [39]. This
hierarchical nanostructure was formed without using any surfactant as well as tem-
plates, and the reaction temperature was at 160 °C for 30 min. Wang et al. reported
the two-step hydrothermal processes for the fabrication of CoO/SnO2 hierarchical
heterojunction nanostructure. The author mentioned the first hydrothermal process
is to design the flowerlike hierarchical SnO2 formed at 180 °C for 12 h, and then the
CoO nanoparticles were developed on the hierarchical SnO2 surfaces in the second
reaction process at the same temperature and reaction time (Fig. 3.3).

3.2.3  Solvothermal Method

In this typical synthesis process, the whole procedure and conditions of the reaction
processes are same as a hydrothermal method, but the only difference is non-­
aqueous solvent, instead of water. The reaction process takes place in an autoclave,
made up of stainless steel with Teflon coating. The pressure and temperature of the
reaction medium can be greater than the hydrothermal process because some
organic solvents used in the solvothermal method have a higher boiling point than
the water. The final synthesized particles have homogenous size distribution with
higher crystalline nature. The main advantages of using the solvothermal method
are controlled morphologies obtained by controlling the temperature and reaction
time [40–42].

Fig. 3.3  Evolution of flowerlike hierarchical ZnO nanostructures by changing the CTAB concen-
tration in the hydrothermal process [77]
3  Hierarchical Nanostructures for Photocatalytic Applications 71

Xiang et al. demonstrated the solvothermal method for the fabrication of titania-­
based hierarchical nanostructure by using TiCl4 and tetrabutyltitanate (TBT) as pre-
cursors and toluene as a solvent [43]. Sea urchin structure was formed at 150 °C for
24 h in the solvothermal reaction medium. The author demonstrated the morpho-
logical formation mechanism was attributed via nucleation-self-assembly-­
dissolution-recrystallization process. Controllable growth and tuneable morphology
of the hierarchical TiO2 nanostructure were demonstrated by Fan et al. [44]. Based
on the microscopic investigation, the author mentioned that the morphology of the
TiO2 can be controlled from rose-like, chrysanthemum-like and sea urchin-like hier-
archical nanostructures by only altering the reaction temperature of the solvother-
mal method (Fig. 3.4).
Xiao et al. published in the literature that the fabrication of flowerlike tungsten
trioxide (WO3) hierarchical nanostructure has been produced via simple one-step
solvothermal process at 100 °C temperature without using any templates and surfac-
tants. The author also conducted the time depended on experimental procedures to
determine the formation mechanism of WO3 nanoflower. It is found that the solvo-
thermal reaction process undergoes a series of growth mechanisms like polymerisa-
tion procedures, nucleation and growth assembly from WO42− to the flowerlike
structure [45].

3.2.4  Microwave Treatment

Microwave treatment provides a facile and rapid synthesis of semiconductor-based


metal oxide hierarchical nanostructures [46]. This synthesis route is comparatively
a unique approach that does not require high temperature and high pressure. The
key feature on this approach is the ability to produce the heat via electromagnetic

Fig. 3.4  Stepwise formation of WO3 hierarchical nanostructure by solvothermal synthesis pro-
cess [78]
72 R. A. Rakkesh et al.

radiation and simultaneously synthesize the material at a very period of time. The
microwave treatment approach works on the basis of two principles, rotation of
dipoles and ionic conduction. The solvents/reactants with high microwave extinc-
tion coefficient can induce the efficient absorption of microwaves and thus signifi-
cantly enhance the reaction rates [47, 48].
There are a huge number of research pieces of literature available on the micro-
wave treatment for the fabrication of hierarchical nanostructures. The microwave
treatment provides an ultrafast decomposition of precursor materials and can syn-
thesize different varieties/forms of metal oxide nanostructures with various size,
morphology and compositions. Thus, in fabrication of semiconducting materials
using microwave treatment, the rate of the reaction can be controlled effectively,
and the resultant material shows homogenous dispersion, and excellent crystalline
nature is formed at a very rapid time. The prime advantages of microwave approach
are high efficiency, rapid processing time and economically viable route for the
fabrication of hierarchical nanostructures [49, 50] (Fig. 3.5).
Yin et al. demonstrated the simple redox reaction processes between tin ions and
nanosheets of graphene oxide to fabricate the SnO2–rGO hierarchical nanostructure
under microwave synthesis approach [51]. The author used in situ heteronucleation
and growth of SnO2 nanocrystals on the surfaces of the graphene sheets under
microwave irradiation for the period of 3  min at 600  W, 2.45  GHz. The author
claimed that the microwave-synthesized hierarchical structure possesses a higher
surface area and shows better activity due to the synergistic effect. Hierarchical ZnO
have been fabricated with mixed solvent conditions (water and ethylene glycol)
through a novel microwave synthesis approach, as demonstrated by Zhu and co-­
workers [52]. The author achieved various morphologies of nanoparticles by only
changing the heating parameters like microwave power, duration and pulse rate of

Fig. 3.5  Scanning electron micrographs of microwave-synthesized zinc oxide nanostructures syn-
thesized at various microwave exposure power: (a) 200 W, (b) 400 W, (f) 600 W and (c–e) magni-
fied images of respective microwave power [79]
3  Hierarchical Nanostructures for Photocatalytic Applications 73

the microwave equipment. It is also proven that the usage of ethylene glycol (EG)
in the reaction medium can act as a reducing agent and stabilizer and prevent the
agglomeration of the particles. EG has high cohesive energy and better dielectric
constant, which makes an excellent microwave radiation acceptor and thus greatly
facilitates the microwave technique for the fabrication of hierarchical
nanostructures.

3.2.5  Metal-Organic Framework (MOF)-Directed Synthesis

MOF-directed synthesis approach is one of the emerging routes for synthesizing


hierarchical nanostructures by forming the coordination bonds between organic
ligands and transition-metal cations. Generally, suitable organic ligands should be
monitored and analysed for the fabrication of hierarchical nanostructures with
transition-­metal cations under elevated conditions. The 2D metal-organic frame-
works (2D MOFs) have received great attention to develop various forms of photo-
catalysts due to their structural engineering capabilities and huge specific surface
area that allows to show a huge number of active sites on the surfaces. Hence, metal-­
organic framework-directed synthesis approach is a newly developed method for
the fabrication of transition metal-based 3D network structure with porous arrange-
ment for energy- and environmental-related fields [53, 54].
Gumilar et al. reported a common approach to fabricate benzene dicarboxylic
acid-supported metal-organic frameworks with 3D hybrid nanostructures composed
of 2D nanosheets [55]. The author used acetonitrile as a solvate for metal ions, and
polyvinylpyrrolidone acts as a morphology controlling mediator to facilitate the
particle growth and nucleation of metal-organic framework nanosheets. The author
demonstrated the flexibility of using several transition-metal ions (M = Cu, Mn, Ni
and Zr) in this synthesis approach. Hao et  al. reported the synthesis of N- and
S-doped C-encapsulated CeO2 hinge-like hierarchical nanostructures has been
obtained by metal-organic framework-directed synthesis approach with Ce-metal-­
organic framework-808 as a precursor. The author reported that the synergistic
effect associated with its hierarchical structure in N- and S-doped C-encapsulated
CeO2 enhances the activity of photocatalytic hydrogen evolution reaction [56]
(Fig. 3.6).
Zinc (II) 2,5-dihydroxy-1,4-benzenedicarboxylate metal-organic frameworks
have been developed by Yue et al. with the combination of microporous-­mesoporous
hierarchical nanostructures without using any templates [57]. The author reported
that the hierarchical nanostructures having nanoporous walls, higher surface area
and larger mesoporous were fabricated within 15 min of reaction at elevated tem-
perature. It is also mentioned that the nanoparticle morphology can be varied by
processing time as well nature of the solvent used. Zhang et al. demonstrated the
metal-organic framework-directed synthesis approach for the fabrication of nano-
structure in a hierarchical manner made up of Fe2O3 nanotubes arranged in Co3O4
host [58]. The author incorporates MIL-88B nanorods in the ZIF-67 polyhedron
74 R. A. Rakkesh et al.

Fig. 3.6  Process flow for the synthesis of Fe2O3 @ Co3O4 hierarchical nanostructures by metal-­
organic framework-directed synthesis method [80]

host to convert the MIL-88B nanorods and ZIF-67 polyhedron to Fe2O3 nanotubes
and Co3O4 host, respectively, by thermal treatment processes. It is also described
that the unique structural arrangements and compositional feature make this mate-
rial as efficient for energy applications.

3.3  Significant Applications of Hierarchical Photocatalysts

In a photocatalytic application, the metal oxide-based semiconductor is a prime


active material, which is able to take responsibilities for both light absorption and
carrier separation properties. As it has been discussed in the synthesis part, the hier-
archical photocatalysts offer a lot of advantages owing to its distinct morphology,
excellent charge separation characteristics and large surface area. Till now, metal
oxide-based hierarchical nanostructures have been utilized in different fields of pho-
tocatalysis. Herein, we have highlighted some of the recent advancement in the
photocatalytic application like photocatalytic water remediation, CO2 reduction and
photocatalytic hydrogen production [59–61].

3.3.1  Photocatalytic Water Remediation

Photocatalytic water remediation is a process of removing hazardous contamina-


tions (organic dyes, pharmaceutical waste and colouring agents) present in the
waterbodies, which remain a hectic trouble in our ecosystem. Industries like leather
processing industries, textile industries, refinery plants, etc. released their harmful
pollutants in the waterbodies, which contain highly toxic chemicals. These harmful
pollutants will demolish our environment in the near future. The photocatalytic per-
formance of metal oxide-based hierarchical nanostructure is greatly useful for the
disinfection of toxic pollutants from the waterbodies [62, 63].
3  Hierarchical Nanostructures for Photocatalytic Applications 75

Ta et al. established the fast photocatalytic remediation of multiple organic dyes


using mixed dimensional Ag-ZnO hierarchical nanostructures [64]. The author
found that 96.5% of dyes were tainted within 30 min under the irradiation of natural
solar light using this hierarchical photocatalyst. It is also examined that the material
showed better recyclability after trials of the same experiment. The author disclosed
the photocatalytic efficiency was based on the rapid electron transfer and surface
plasmon resonance mechanism between the hierarchical photocatalysts and the pol-
lutants. New ordered nanostructured arrays made up of Ag–ZnO hierarchical nano-
structure were synthesized by Liu and co-workers for the water remediation of
Congo red pollutant and bacterial species in the waterbodies [65]. The authors
investigated the photocatalytic activity of ZnO–Ag hierarchical nanoarrays for the
disinfection of Congo red dye, as it suppresses the recombination rate of charge car-
riers and also broadens the absorption band from ultraviolet to the visible range. The
author found that the fabricated photocatalyst exhibits highest photocatalytic degra-
dation efficiency and potent material to kill the bacterial species in the waterbodies
effectively at very short duration without producing any adverse effect on the envi-
ronment (Fig. 3.7).
Zhang et al. investigated the enhanced photocatalytic performance of copper (II)
phthalocyanine on electrospun TiO2 nanofibers (TNCuPc/TiO2) [66]. The author
showed the photocatalytic activity of hierarchical TNCuPc/TiO2 nanostructures dis-
played improved activity for degradation of rhodamine B dye. Moreover, hierarchi-
cal nanostructures of TNCuPC/TiO2 showed better recyclable capacity without any
change in the photocatalytic performances due to their excellent structural arrange-
ment. It is also expected by the author that the very high photocatalytic activity of

Fig. 3.7  Charge transfer mechanism of Ag–ZnO hierarchical nanostructure-based photocata-


lysts [81]
76 R. A. Rakkesh et al.

hierarchical nanostructured TNCuPC/TiO2 will significantly encourage their com-


mercial application to degrade the toxic contaminants from wastewater. Malwal
et al. reported the fabrication of zinc oxide and iron oxide was placed on the array
surfaces of CuO nanowire for photocatalytic applications [67].
In the relative performance analysis, the author found the arrangement of p-n
heterojunction at the boundary of p-type (CuO) and n-type semiconductor (zinc
oxide and iron oxide), which hinders the electron–hole pair recombination rate [68].
More interestingly, the author also performed the photocatalytic activity of the syn-
thesized hierarchical nanostructure in both extremely acidic and basic conditions to
determine the stability and repeatability of degradation ability. In the perspective of
commercialization, such hierarchical nanostructures are more effective due to its
distinct foamy structure, which helps prevent and degrade the pollutants in the water
system [69] (Fig. 3.8).

Fig. 3.8  Photocatalytic degradation data plotted between concentrations of the pollutant and irra-
diation time [82]
3  Hierarchical Nanostructures for Photocatalytic Applications 77

3.3.2  Photocatalytic CO2 Reduction

Photocatalytic CO2 reduction is a chemical process whereby carbon dioxide is


reduced to hydrocarbon by the energy of incident light. It is one of the greenhouse
gases and contributes to climate change issues in our environment. The increase in
the level of CO2 emission produces drastic climate change, thus causing global
warming problem. Converting CO2 into a useful renewable fuel would be a proper
method of treating shortage of energy fuels as well as global warming issues. Hence,
this will be a fascinating research field of utilizing fossil fuel as renewable energy,
which requires appropriate catalyst and energy input. Therefore, novels design for
sustainable development of hierarchical nanostructure-based photocatalyst for the
reduction of CO2 into useful energy fuel with the help of solar light [70–72].
Wang and his co-workers designed a sandwich-like hierarchical nanostructure of
ZnIn2S4-In2O3 in tubular arrangement; it was assembled like microtubule structure
of In2O3 and was covered both inner and outer surfaces by growing ZnIn2S4
nanosheets [73]. The author demonstrated the photocatalytic performances of dou-
ble heterojunction photocatalyst and found it to be an efficient material for CO2
reduction. The unique structural arrangement promotes the photogenerated charge
carrier transfer and separation effectively, thus enabling adsorption of CO2 in the
large surface area and allowing surface catalytic reaction in the exposed active sites
of the hierarchical structure. It is reported that the potent hierarchical ZnIn2S4–In2O3
nanostructure portrayed excellent activity for the reduction of CO2 with consider-
able CO evolution rate (3075 μmol h−1 g−1) and better stability [74] (Fig. 3.9).

Fig. 3.9  Electron micrographs of sandwich-like hierarchical nanostructure of ZnIn2S4–In2O3 in a


tubular arrangement [83]
78 R. A. Rakkesh et al.

Similarly, Ziarati et al. fabricated the urchin-like yolk@shell TiO2−xHx hierarchi-


cal nanoparticles decorated with plasmonic Au–Pd core/shell nanoparticles
(HUY@S-TOH/AuPd) and demonstrated the photocatalytic CO2 conversion activity
under visible light irradiation [75]. The author reported the obtained yolk@shell
urchin-like structure has rich oxygen vacancies and also plentiful needle-like shapes
decorated with plasmonic Au–Pd nanoparticles over the hierarchical arrangement.
The reported nanostructure revealed excellent CO2 to CH4 values of up to
47  μmol  g−1  h−1 which signifies excellent photocatalytic performance over the
recently reported studies. The author explained the high photocatalytic reduction per-
formance mainly attributed by the mixture of band position, surface defects, porosity
and co-catalyst and facilitates a proficient adsorption capacity of the target gas, better
charge carrier recombination and transport to influential CO2 photoconversion [76].
Jung et al. developed the extremely stable and proficient hierarchical photocata-
lysts for the reduction of CO2 using TiO2 mesoporous structures resting on graphene
3D matrix with few layers of MoS2 nanostructures [77]. The author investigated the
performance of the nanostructures by controlling the morphologies and thus enhanc-
ing the performance, robustness and stable CO2 reduction. The reported porous
hierarchical nanostructure shows better nanoarchitecture with economically viable
route, robust technology and higher catalytic efficiency for the conversion of CO2 to
selective CO (Fig. 3.10).

Fig. 3.10  CO formation rate of TiO2 mesoporous structures resting on graphene 3D matrix with
few layers of MoS2 nanostructures [84]
3  Hierarchical Nanostructures for Photocatalytic Applications 79

3.3.3  Photocatalytic H2 Production

Hierarchical nanostructure for photocatalytic H2 production through water splitting


produces an economically friendly, clean and effective consumption of abundant
energy from sunlight, which can be effectively utilized in the field of electric devices
and vehicles. The visible light- or sunlight-directed photocatalytic water splitting
reaction involves the following chemical reactions: (i) hydrogen (H2) and oxygen
(O2) are separated from the water molecules by photoinduced chemical reduction
reactions occurring at the semiconductor (photocatalysis) surfaces. (ii) The photo-
generated charge carrier (electrons) reacts with H+ generated from water molecules
and undergoes in situ reduction into to H2, while the photoinduced holes oxidize
H2O to H+ and O2. In such a way, hydrogen fuel was produced and stored by using
sunlight active photocatalytic water splitting or reforming processes [78–80]
(Fig. 3.11).
The core-shell ZnO–ZnS–Cu2S-based hierarchical arrayed nanostructures were
fabricated by Ranjith and co-workers [81]. The author demonstrated the fabricated
nanostructured photocatalyst operated under visible light active hydrogen evolution
reaction. It is found to be 436 μmol h−1 g−1 irradiated under visible light. The author
claimed that the enhanced effectiveness is accredited to the effectual Cu2S–ZnS
heterojunction on the zinc oxide nanorods, which endorses better separation effi-
ciency of charge carriers by Z scheme photocatalytic processes. Li et al. developed

Fig. 3.11 (a) SEM images, (b) valence band structures and (c) electron transport mechanism of
ZnO/ZnS hierarchical nanostructures [85]
80 R. A. Rakkesh et al.

the featherlike orientation of TiO2 catalyst for efficient photocatalytic hydrogen pro-
duction [82]. Such hierarchical nanostructure has a porous texture with branched
pattern facilitates for the very large surface area associated with more active sites for
photocatalytic reaction. The author investigated the visible-light active photocata-
lytic performances which were achieved at 44.74 mmol m−2 h−1 rates of hydrogen
production, which is far higher than the recently reported works of literature. The
photocatalytic water splitting activity remains the same after 100 h cycle test, which
demonstrates the stability of the fabricated hierarchical nanostructures. The author
highlighted this structure as an efficient hierarchical structure for better photocata-
lytic performance water splitting application [83] (Fig. 3.12).
Si et al. presented their research work on hierarchical macro-mesoporous poly-
meric carbon nitride microspheres with narrow bandgap for excellent photocatalytic
hydrogen production [84]. The reported architecture has a unique hierarchical struc-
ture with the combination of both macro-mesoporous nanostructure. This hierarchi-
cal nanostructure provides bandgap energy, which facilitates a wide range of light
absorption and enables excellent charge transportation and separation processes
[85]. The author demonstrated the hierarchical structure of macro-mesoporous
polymeric carbon nitride microspheres shows excellent hydrogen production

Fig. 3.12  Hydrogen and oxygen evolution profile from photocatalytic water splitting pro-
cesses [86]
3  Hierarchical Nanostructures for Photocatalytic Applications 81

performance (4635.8  μmol  h−1  g−1). The unique hierarchical macro-mesoporous


nanostructure for getting better solar energy capture and conversion generates novel
viewpoints for fabricating rigid and stable photocatalysts for sunlight active photo-
catalytic hydrogen production application in the near future.

3.4  Conclusions

This chapter reviewed current research progress in the basic concepts, synthesis
strategies and application of hierarchical nanostructure-based photocatalysts. These
materials demonstrated higher photocatalytic activity for degradation of pollutants;
thus, it paves the way to solve the energy- and environmental-related problems.
These materials are successful due to its very high specific surface area, exposed
crystal facets, unique morphology, great absorption property and better charge
transportation and separation efficiencies. Among the various nanomaterial sys-
tems, metal oxide-based semiconductors have been utilized to examine the suitable
hierarchical structures. This chapter reviewed the basic concepts of hierarchical
nanostructures, fabrication strategies and significant applications related to energy
and environmental fields. Particularly this chapter clearly elucidates the novel syn-
thesis techniques such as precipitation synthesis, hydrothermal method, solvother-
mal method, microwave treatment and metal-organic framework-directed synthesis
approach for the synthesis of metal oxide semiconducting-based hierarchical nano-
structures. In the end, some key applications like photocatalytic water remediation,
photocatalytic carbon dioxide and photocatalytic hydrogen fuel production have
been reviewed clearly.

References

1. Li, X., Wen, J., Low, J., Fang, Y., & Yu, J. (2014). Science China Materials, 57, 70–100. https://
doi.org/10.1007/s40843-­014-­0003-­1.
2. Li, X., Yu, J., Low, J., Fang, Y., Xiao, J., & Chen, X. (2015). Journal of Materials Chemistry A,
3, 2485–2534. https://doi.org/10.1039/C4TA04461D.
3. Bard, & Fox, M. (1995). Accounts of Chemical Research, 28, 141–145. https://doi.org/10.1021/
ar00051a007.
4. Hoffmann, M., Martin, S., Choi, W., & Bahnemann, D. (1995). Chemical Reviews, 95, 69–96.
https://doi.org/10.1021/cr00033a004.
5. Kudo, & Miseki, Y. (2009). Chemical Society Reviews, 38, 253–278. https://doi.org/10.1039/
B800489G.
6. R.  Ajay Rakkesh, D.  Durgalakshmi & S.  Balakumar (2017) John Wiley and Scrivener
Publishing, USA (ISBN: 978-1-119-16034-2).
7. R.  Ajay Rakkesh, D.  Durgalakshmi & S.  Balakumar (2017) Springer-Nature, USA (ISBN:
978-3-319-62446-4).
8. Chen, X., Shen, S., Guo, L., & Mao, S. S. (2010). Chemical Reviews, 110, 6503–6570. https://
doi.org/10.1021/cr1001645.
82 R. A. Rakkesh et al.

9. Lang, X., Chen, X., & Zhao, J. (2014). Chemical Society Reviews, 43, 473–486. https://doi.
org/10.1039/C3CS60188A.
10. Fujishima, A., & Honda, K. (1972). Nature, 238, 37–38. https://doi.org/10.1038/238037a0.
11. Rajeshwar, K. (2007). Journal of Applied Electrochemistry, 37, 765–787. https://doi.

org/10.1007/s10800-­007-­9333-­1.
12. Maeda, K. (2011). Journal of Photochemistry and Photobiology C, 12, 237–268. https://doi.
org/10.1016/j.jphotochemrev.2011.07.001.
13. Ajay Rakkesh, R., Durgalakshmi, D., Karthe, P., & Balakumar, S. (2020). Materials Chemistry
and Physics. https://doi.org/10.1016/j.matchemphys.2020.122720.
14. Ajay Rakkesh, R., Durgalakshmi, D., Karthe, P., & Balakumar, S. (2019). Journal of Processing
and Application of Ceramics, 13, 376–386. https://doi.org/10.2298/PAC1904376A.
15. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2018). ChemistrySelect, 3, 7302–7309.
https://doi.org/10.1002/slct.201800987.
16. Hisatomi, T., Kubota, J., & Domen, K. (2014). Chemical Society Reviews, 43, 7520–7535.
https://doi.org/10.1039/C3CS60378D.
17. Moriya, Y., Takata, T., & Domen, K. (2013). Coordination Chemistry Reviews, 257,

1957–1969. https://doi.org/10.1016/j.ccr.2013.01.021.
18. Pelaez, M., Nolan, N. T., Pillai, S. C., Seery, M. K., Falaras, P., Kontos, A. G., Dunlop, P. S. M.,
Hamilton, J. W. J., Byrne, J. A., O’Shea, K., Entezari, M. H., & Dionysiou, D. D. (2012). Applied
Catalysis B: Environmental, 125, 331–349. https://doi.org/10.1016/j.apcatb.2012.05.036.
19. Di Paola, E., Garcia-Lopez, G., Marci, G., & Palmisano, L. (2012). Journal of Hazardous
Materials, 211, 3–292.
20. Martha, S., Sahoo, P. C., & Parida, K. M. (2015). RSC Advances, 5, 61535–61553. https://doi.
org/10.1039/C5RA11682A.
21. Rani, R., Reddy, U., Sharma, P., Mukerjee, P., Mishra, A. K., & Sim, L. C. (2018). Journal of
Nanostructure in Chemistry, 8, 255–291.
22. Li, H., Zhou, Y., Tu, W., Ye, J., & Zou, Z. (2015). Advanced Functional Materials, 25,
998–1013. https://doi.org/10.1002/adfm.201401636.
23. Li, H., Tu, W., Zhou, Y., & Zou, Z. (2016). Advancement of Science, 3, 1500389.
24. Ohtani, B. (2010). Journal of Photochemistry and Photobiology C Photochemistry Reviews,
11, 157–178. https://doi.org/10.1016/j.jphotochemrev.2011.02.001.
25. Pallavi, N., & Shivaraju, H. P. (2017). International Journal of Nanotechnology, 14, 762–774.
https://doi.org/10.1504/IJNT.2017.086762.
26. Lewis, N. S., & Nocera, D. G. (2006). Proceedings of the National Academy of Sciences of the
United States of America, 103, 15729. https://doi.org/10.1073/pnas.0603395103.
27. Walter, M. G., Warren, E. L., McKone, J. R., Boettcher, S. W., Mi, Q., Santori, E. A., & Lewis,
N. S. (2010). Chemical Reviews, 110, 6446. https://doi.org/10.1021/cr1002326.
28. Gasteiger, H. A., Kocha, S. S., Sompalli, B., & Wagner, F. T. (2005). Applied Catalysis B:
Environmental, 56, 9. https://doi.org/10.1016/j.apcatb.2004.06.021.
29. Gasteiger, H.  A., & Marković, N.  M. (2009). Science, 324, 48. https://doi.org/10.1126/

science.1172083.
30. Osterloh, F.  E. (2013). Chemical Society Reviews, 42, 2294. https://doi.org/10.1039/

C2CS35266D.
31. Wang, Z. L. (2004). Materials Today, 7, 26–33. https://doi.org/10.1016/S1369-­7021(04)00286-­X.
32. Xu, S., & Wang, Z.  L. (2011). Nano Research, 4, 1013–1098. https://doi.org/10.1007/

s12274-­011-­0160-­7.
33. Burke-Govey, C. P., & Plank, N. O. V. (2013). Journal of Vacuum Science and Technology B,
31, 06F101. https://doi.org/10.1116/1.4821801.
34. Sun, H., Yu, Y., Luo, J., Ahmad, M., & Zhu, J. (2012). CrystEngComm, 14, 8626–8632. https://
doi.org/10.1039/c2ce26157j.
35. Alenezi, M. R., Henley, S. J., Emerson, N. G., & Silva, S. R. (2014). Nanoscale, 6, 235–247.
https://doi.org/10.1039/C3NR04519F.
3  Hierarchical Nanostructures for Photocatalytic Applications 83

36. Liu, X., Zhao, J., Cao, Y., Li, W., Sun, Y., Lu, J., Men, Y., & Hu, J. (2015). RSC Advances, 5,
47506–47510. https://doi.org/10.1039/C5RA05231A.
37. Ma, Y., Bian, Y., Liu, Y., Zhu, A., Wu, H., Cui, H., Chu, D., & Pan, J. (2018). ACS Sustainable
Chemistry & Engineering, 6, 2552–2562.
38. Ray, C., & Pal, T. (2017). Journal of Materials Chemistry A, 5, 9465–9487; Li, J., Zhang,
M., Li, X., Li, Q., & Yang (2017). Journal of Applied Catalysis B, 212, 106–114. https://doi.
org/10.1039/C7TA02116J.
39. Zou, X., & Zhang, Y. (2015). Chemical Society Reviews, 44, 5148–5180. https://doi.

org/10.1039/C4CS00448E.
40. Zhang, H., Lv, X., Li, Y., Wang, Y., & Li, J. (2010). ACS Nano, 4, 380–386.
41. Alivisatos, P. (1996). Science, 271, 933–937.
42. Hu, X., Yu, J. C., Gong, J., & Li, Q. (2007). Crystal Growth & Design, 7, 2444–2448. https://
doi.org/10.1021/cg060767o.
43. Ho, W., Yu, J.  C., & Lee, S. (2006). Chemical Communications, 1115–1117. https://doi.
org/10.1039/b515513d.
44. Sing, K., Everett, D., Haul, R., Moscou, L., Pierotti, R., Rouquerol, J., & Siemieniewska, T.
(1985). Pure and Applied Chemistry, 57, 603–619. https://doi.org/10.1351/pac198557040603.
45. Rolison, D. R. (2003). Science, 299, 1698–1701. https://doi.org/10.1126/science.1082332.
46. Zhou, W., & Fu, H. (2013). ChemCatChem, 5, 885–894. https://doi.org/10.1002/

cctc.201200519.
47. Li, G., Zhang, D., & Yu, J.  C. (2008). Chemistry of Materials, 20, 3983–3992. https://doi.
org/10.1021/cm800236z.
48. Du, J., Lai, X., Yang, N., Zhai, J., Kisailus, D., Su, F., Wang, D., & Jiang, L. (2011). ACS Nano,
5, 590–596.
49. X. C. Wang, J. C. Yu, C. M. Ho, Y. D. Hou & X. Z. Fu, Langmuir, 2005, 21, 2552–2559, DOI:
10.1021/la047979c.
50. Xi, G.  C., & Ye, J.  H. (2010). Chemistry: A European Journal, 16, 8719–8725. https://doi.
org/10.1002/chem.200903380.
51. Yu, J., Su, Y., & Cheng, B. (2007). Advanced Functional Materials, 17, 1984–1990. https://
doi.org/10.1002/adfm.200600933.
52. Yuan, Z. Y., Ren, T. Z., & Su, B. L. (2003). Advanced Materials, 15, 1462–1465. https://doi.
org/10.1002/adma.200305075.
53. Cheng, C., & Fan, H.  J. (2012). Nano Today, 7, 327–343. https://doi.org/10.1016/j.

nantod.2012.06.002.
54. Liu, S., Yang, M.-Q., Tang, Z.-R., & Xu, Y.-J. (2014). Nanoscale, 6, 7193–7198. https://doi.
org/10.1039/c4nr01227e.
55. Xiao, F.-X., Hung, S.-F., Tao, H. B., Miao, J., Yang, H. B., & Liu, B. (2014). Nanoscale, 6,
14950–14961. https://doi.org/10.1039/C4NR04886E.
56. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2016). RSC Advances, 6, 34342–34349.
https://doi.org/10.1039/C6RA01784C.
57. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2015). RSC Advances, 5, 18633–18641.
https://doi.org/10.1039/C5RA00180C.
58. Zhao, Y., Wang, W., Li, Y., Zhang, Y., Yan, Z., & Huo, Z. (2014). Nanoscale, 6, 195–198.
https://doi.org/10.1039/C3NR04280D.
59. Jiao, Y., Liu, Y., Qu, F., & Wu, X. (2014). CrystEngComm, 16, 575–580. https://doi.

org/10.1039/C3CE41994K.
60. Xiang, Q., Yu, J., & Jaroniec, M. (2011). Chemical Communications, 47, 4532–4534. https://
doi.org/10.1039/c1cc10501a.
61. Athauda, T. J., Neff, J. G., Sutherlin, L., Butt, U., & Ozer, R. R. (2012). ACS Applied Materials
& Interfaces, 4, 6916–6925.
62. Han, C., Chen, Z., Zhang, N., Colmenares, J. C., & Xu, Y.-J. (2015). Advanced Functional
Materials, 25, 221–229. https://doi.org/10.1002/adfm.201402443.
84 R. A. Rakkesh et al.

63. D.  Chatterjee & S.  Dasgupta, Journal of Photochemistry and Photobiology C, 2005, 6,
186–205, DOI: 10.1016/j.jphotochemrev.2005.09.001.
64. Xu, D., Cheng, B., Cao, S., & Yu, J. (2015). Applied Catalysis B: Environmental, 164, 380–388.
https://doi.org/10.1016/j.apcatb.2014.09.051.
65. Liu, Y., Wang, R., Yang, Z., Du, H., Jiang, Y., Shen, C., Liang, K., & Xu, A. (2015). Chinese
Journal of Catalysis, 36, 2135–2144. https://doi.org/10.1016/S1872-­2067(15)60985-­8.
66. Xiong, T., Zhang, H., Zhang, Y., & Dong, F. (2015). Chinese Journal of Catalysis, 36,
2155–2163. https://doi.org/10.1016/S1872-­2067(15)60980-­9.
67. Yu, C., Bai, Y., He, H., Fan, W., Zhu, L., & Zhou, W. (2015). Chinese Journal of Catalysis, 36,
2178–2185. https://doi.org/10.1016/S1872-­2067(15)61009-­9.
68. Wang, X., Yu, R., Wang, K., Yang, G., & Yu, H. (2015). Chinese Journal of Catalysis, 36,
2211–2218.
69. Zhang, X., Zhu, Y., Yang, X., Zhou, Y., Yao, Y., & Li, C. (2014). Nanoscale, 6, 5971–5979.
https://doi.org/10.1039/C4NR00975D.
70. Ai, Z. H., Zhang, L. Z., Lee, S. C., & Ho, W. K. (2009). Journal of Physical Chemistry C, 113,
20896–20902. https://doi.org/10.1021/jp9083647.
71. Ma, B., Guo, J., Dai, W.-L., & Fan, K. (2012). Applied Catalysis B: Environmental, 123,
193–199.
72. Shen, Z., Chen, G., Wang, Q., Yu, Y., Zhou, C., & Wang, Y. (2012). Nanoscale, 4, 2010–2017.
https://doi.org/10.1039/c2nr12045c.
73. Sun, M., Yan, Q., Yan, T., Li, M., Wei, D., Wang, Z., Wei, Q., & Du, B. (2014). RSC Advances,
4, 31019–31027. https://doi.org/10.1039/C4RA03843F.
74. Xing, C., Zhang, Y., Wu, Z., Jiang, D., & Chen, M. (2014). Dalton Transactions, 43, 2772–2780.
https://doi.org/10.1039/C3DT52875H.
75. Dong, F., Sun, Y., Fu, M., Wu, Z., & Lee, S.  C. (2012). Journal of Hazardous Materials,
219, 26–34.
76. Guzman, S.  S., Jayan, B.  R., de la Rosa, E., Castro, A.  T., Gonalez, V.  G., & Yacaman,
M.  J. (2009). Materials Chemistry and Physics, 115, 172–178. https://doi.org/10.1016/j.
matchemphys.2008.11.030.
77. Zhu, L., Li, Y., & Zeng, W. (2017). Applied Surface Science, 427, 281–287.
78. Xiao, B., Zhao, Q., Xiao, C., Yang, T., Wang, P., Wang, F., Chen, X., & Zhang, M. (2015).
CrystEngComm. https://doi.org/10.1039/C5CE00870K.
79. Zhu, P., Zhang, J., Wu, Z., & Zhang, Z. (2008). Crystal Growth and Design, 8, 3148–3153.
https://doi.org/10.1021/cg0704504.
80. Zhang, S. L., Guan, B. Y., Wu, H. B., & Lou, X. W. D. (2018). Nano-Micro Letters, 10, 44.
81. Liu, J., Li, J., Wei, F., Zhao, X., Su, Y., & Han, X. (2019). ACS Sustainable Chemistry &
Engineering, 7, 11258–11266. https://doi.org/10.1021/acssuschemeng.9b00610.
82. Zhang, M., Shao, C., Guo, Z., Zhang, Z., Mu, J., Cao, T., & Liu, Y. (2011). ACS Applied
Materials & Interfaces, 3, 369–377. https://doi.org/10.1021/am100989a.
83. Wang, S., Guan, B. Y., & Wen, X. (2018). Journal of the American Chemical Society. https://
doi.org/10.1021/jacs.8b02200.
84. Jung, H., Cho, K. M., Kim, K. H., Yoo, H. W., Saggaf, A. A., Gereige, I., & Jung, H. T. (2018).
ACS Sustainable Chemistry & Engineering. https://doi.org/10.1021/acssuschemeng.8b00002.
85. K. S. Ranjith, D. R. Kumar, Y. S. Huh, Y. K. Han, T. Uyar & R. T. R. Kumar (2020). https://doi.
org/10.1021/acs.jpcc.9b09666.
86. Li, G., Huang, J., Xue, C., Chen, J., Deng, Z., Huang, Q., Liu, Z., Gong, C., Guo, W., & Cao,
R. (2019). Crystal Growth & Design. https://doi.org/10.1021/acs.cgd.9b00495.
87. Ajay Rakkesh, R., & Balakumar, S. (2015). Journal of Nanoscience and Nanotechnology, 15,
4316–4324. https://doi.org/10.1166/jnn.2015.9723.
88. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2015). AIP Conference Proceedings,
1665, 050036.
89. Ajay Rakkesh, R., Durgalakshmi, D., & Balakumar, S. (2014). Journal of Materials Chemistry
C, 2, 6827–6834. https://doi.org/10.1039/C4TC01195C.
Chapter 4
Nanocomposite Photocatalysts
for the Degradation of Contaminants
of Emerging Concerns

Rokesh Karuppannan, Sakar Mohan, and Trong-On Do

4.1  Introduction

Antibiotics is one of the most prescribed and consumed pharmaceutical drugs


widely used for bacterial diseases. They play critical roles in causing some adverse
effects to the human health and ecosystem when they get released to the environ-
ment without proper safety measures. These pharmaceutical drugs primarily con-
taminate the water when they get released into water bodies. The continuous release
of antibiotics into aquatic environment promotes bacterial resistivity and generates
stable organic by-products. These antibiotic resistances can essentially cause geno-
toxic effects, which could mutate the bacteria and possibly generate “superbugs” to
human society. Similarly, the antibiotic residues also generate some eco-toxic
effects to aquatic animals. Hence, the presence of antibiotic residue molecules in
wastewater and their harmful effects on ecosystems and human life have received
huge consideration worldwide; therefore, these antibiotics are considered as “con-
taminants of emerging concerns” and “emerging pharmaceutical pollutants” [1–3].
Hence, there is an urgent need to address these emerging concerns and to develop
an efficient and sustainable technique for the degradation of antibiotic pollutants
from the aqueous systems. There are numerous techniques available to remove anti-
biotic chemicals from water, which include adsorption, photocatalysis, ozonolysis,
filtration, and biological process as depicted in Scheme 4.1 [4–9]. Among these
techniques, the nanocomposite-based semiconductor photocatalytic technique
unveils a great consideration because of their potential redox ability and superior

R. Karuppannan · T.-O. Do (*)


Department of Chemical Engineering, Laval University, Quebec, QC, Canada
e-mail: Trong-On.Do@gch.ulaval.ca
S. Mohan
Department of Chemical Engineering, Laval University, Quebec, QC, Canada
Centre for Nano and Material Sciences, Jain University, Bangalore, Karnataka, India

© Springer Nature Switzerland AG 2021 85


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_4
86 R. Karuppannan et al.

Scheme 4.1  Various processes for the degradation/removal of pharmaceutical pollutants from
water [4–9]

degradation performance, higher mineralization efficiency, and less/harmless prod-


uct generation [10, 11].
To this end, this chapter presents an overview on the design and development of
nanocomposite photocatalyst systems such as semiconductor-metal, semiconductor-­
semiconductor, and semiconductor-carbon composites and their mechanisms
towards the degradation of antibiotic pollutants in water. It is found that the mecha-
nism of these nanocomposite materials is superior towards their photocatalytic deg-
radations of pharmaceutical pollutants, thanks to their improved photoabsorbance,
charge separation, charge transfer, redox potential, and photostability.

4.2  Overview of Photocatalysis

Photocatalysis has received great attention in antibiotic degradation because it uti-


lizes light energy to perform the potential redox reactions to degrade them. The
photocatalytic degradation reaction can be established by two different reaction
pathways; one is primary and another one is secondary reaction. The primary reac-
tion involves the direct oxidization of the antibiotic molecules by photogenerated
holes (h+). The secondary reaction involves generation of reactive-free radicals by
the oxidization of water molecules with photogenerated holes to produce hydroxyl
radicals (˙OH), where they consequently initiate the oxidation reaction. Furthermore,
the photogenerated electron can be entrapped by oxygen molecule, which leads to
the generation of superoxide radicals (O2˙−) and also oxidizes the antibiotics.
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 87

Among them, the hydroxyl radicals have attracted more attention in photocatalytic
degradation process due to their strong oxidation potential, and also they offer non-
selective destruction, which will be able to interact with a wide range of contami-
nants without any additives. Therefore, the photocatalytic redox reaction with free
radical generation can oxidize and completely mineralize the antibiotic molecules
[12, 13].

4.3  Semiconductor Photocatalysts

Photocatalytic semiconductors are key tools in photocatalytic process. Generally,


the photocatalytic materials undergo the following principal processes such as (i)
photo-absorption, (ii) charge separation, and (iii) charge transfer. Typically, most of
the semiconductor photocatalysts are only ultraviolet (UV) active due to its broad
bandgap and can utilize limited solar energy. On other hand, the narrow band semi-
conductor photocatalyst shows lower photoinduced electron-hole charge separation
and short photogenerated electrons lifetimes, which noticeably encourages the
charge recombination and thereby affects the photocatalytic efficiency. Similarly,
the photocatalytic semiconductor undergoes the surface or bulk recombination and
also involves in photo-corrosion, which eventually limits the stability of the semi-
conductor photocatalyst. Thus, the abovementioned major limitations potentially
reduce the photocatalytic performances of the semiconductor photocatalysts. In this
direction, the development of nanocomposite semiconductor photocatalyst offers
extended photo-absorption, higher charge separation, reduced charge recombina-
tion, and higher photostability [13, 14]. Therefore, constructing the nanocomposite
semiconductor photocatalysts is a promising tool to overcome the said limitations
of the conventional semiconductor photocatalysts.

4.4  Nanostructured Photocatalyst

Nanostructure fabrication offers a potential route for higher photocatalytic perfor-


mance of semiconductor photocatalyst. This is due to the fact that the nanostruc-
tures provide smaller size, high surface areas, various morphologies, and easy
fabrication. The designing of nanostructured photocatalysts shows surface-­
dependent photocatalytic activity as they provide high specific surface area as well
as superior surface reactive sites. Also, the nanostructured photocatalysts facilitate
fast charge carrier migration due to the shorter core to surface distances as com-
pared to the bulk photocatalyst. However, their fast and efficient charge transfer
mechanism is closely correlated with morphology, structure, and surface properties
of the photocatalysts. Hence, all these properties are significantly influencing and
beneficial for the superior photocatalytic activity of nanostructured semiconductor
photocatalyst [13, 15]. For example, the mesoporous titanium dioxide (TiO2)
88 R. Karuppannan et al.

Fig. 4.1  Performance of mesoporous TiO2 nanoaggregates via (a) adsorption process and (b)
photocatalytic process [16]

nanoaggregates were synthesized using titanium glycolate via three different post-
treatment routes such as hydrothermal, calcination, and hydrothermal-calcination
techniques. These posttreatment techniques obviously influenced the TiO2 nanoag-
gregates and their structural formation, and thereby it influenced their catalytic per-
formance as well. Accordingly, the as-developed mesoporous TiO2 nanoaggregates
were studied for their adsorption- and photocatalysis-mediated performances
towards the removal of ciprofloxacin antibiotics (Fig. 4.1) [16].
Similarly, three-dimensional hierarchical architecture of BiOCl provided an
enhanced specific surface area along with efficient electron-hole pair charge separa-
tion. The more positive valence band potential with optimized hierarchical porous
structure has distributed sufficient amount of hydroxyl radicals. Therefore, BiOCl
hierarchical porous materials were found to show superior photocatalytic activities,
and they also showed excellent photocatalytic activity that led to the complete deg-
radation of tetracycline (TC) within 20 min, which was 20 times higher than pure
BiOCl [17]. Then, it was found that the polybutadiene-block-poly(2-vinylpyridine)
cylindrical polymer brushes as template-assisted zinc oxide nanotubes potentially
degraded the ciprofloxacin antibiotics under solar light. The diblock copolymer
template offered non-woven microstructure ZnO nanotubes, which showed 2.9
times greater photocatalytic activity than the Degussa P25 [18].

4.5  Nanocomposite Photocatalyst

The nanocomposite materials show great attention in photocatalytic application due


to their improved charge separation, higher charge transfer, separate reactive sites,
extended photo-absorption and greater photostability. The suitable materials and
their efficient surface and structural properties are determining the factors for the
photophysical and photochemical properties of nanocomposite photocatalyst. The
designing of nanocomposite photocatalysts by combining the nanostructured
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 89

photocatalyst with semiconductor or metal or carbon nanomaterials potentially


overcomes the limitation of nanostructured photocatalysts such as limited solar
light absorption, poor charge separation, higher rate of charge recombination, low
charge transfer efficiency, and photo-corrosion [13, 15]. Therefore, the development
of nanocomposite materials appears to be largely interested in the designing of pho-
tocatalytic materials. Notably, many nanocomposite systems have been demon-
strated for the photocatalytic destruction of antibiotics.

4.5.1  Semiconductor–Metal Composites

The semiconductor–metal composite is extensively studied in photocatalysis


because of their superior photocatalytic performance. The semiconductor-metal
composite offers greater charge separation at the semiconductor-metal interface by
the formation of Schottky barrier, which essentially increases the visible light
absorption through the surface plasmon resonance effect [19, 20].

Schottky Barrier Composites

In the semiconductor–metal-based composites, the Schottky barrier is established


due to the work function difference between the metals and semiconductor. Further,
the Schottky barrier generates an integral electric field near the semiconductor–
metal interface via Fermi level equilibrium, which favors the photogenerated charge
carrier separation and photocatalytic activity [19]. For instance, the bimetallic Ag/
Pt@TiO2 composite showed much higher photocatalytic performance on antibiotic
ciprofloxacin degradation (~100% in 150 min) as compared to Ag@TiO2 and Pt@
TiO2. Due to the formation of Schottky barrier and their synergistic effects between
the metals (Ag and Pt) and TiO2 semiconductor interface, it drastically improved the
charge separation and diffusion. Besides, the incorporation of bimetallic nanopar-
ticles into TiO2 surface was more prominent; therefore it noticeably enhanced the
photocatalytic activity as compared to the single metal incorporation [21]. Similarly,
the Schottky barrier formation between Pt and CdS enhanced the photoinduced
electron-hole pair separation and inhibited the reverse electron migration, thereby
improving the photocatalytic degradation of tetracycline [22]. Likewise, the transi-
tion metal-semiconductor Ag3PO4-NPs/Cu-NWs composite photocatalyst also
favored for the enhanced photocatalytic activity, where the Cu acted as an electron
sink to trap the electrons; thereby the photogenerated free electrons were potentially
transferred from Ag3PO4 to single-crystal Cu nanowires and promoted photogene-
rated electron-hole pair separation (Fig. 4.2). The catalytic performance of Ag3PO4-­
NPs/Cu-NWs was tested on the degradation of antibiotic ciprofloxacin under visible
light irradiation and founded 6.07 times higher degradation efficiency than that of
the pure Ag3PO4 NPs [23].
90 R. Karuppannan et al.

Fig. 4.2  The Ag3PO4-NPs/


Cu-NWs metal-­
semiconductor
nanocomposite
photocatalyst charge
separation and degradation
mechanism [23]

Interestingly, the metal carbides also presented promising metal-like properties


and superior photocatalytic performance. For instance, the designing of Ag3PO4/
Ti3C2 Schottky photocatalyst potentially improved the photocatalytic activity as
well as the photostability of Ag3PO4 towards the degradation of tetracycline.
Because of the strong interfacial contact and different work functions between Ti3C2
and Ag3PO4, they established a strong Schottky junction, and thereby it improved
the charge carrier separation. Notably, the strong redox ability of the surface tita-
nium sites offered the multiple electron reduction and encouraged the rapid hydroxyl
radical production [24].

Plasmonic Composites

In general, the metal nanoparticles can decrease the recombination rate of the pho-
toinduced electron-hole pairs of semiconductor material by effective electron trap-
ping. Correspondingly, the noble metals show surface plasmon resonance (SPR)
effects, which offer broad and strong absorption in the visible region. In this direc-
tion, the gold and silver metals are widely explored plasmonic metal, and their sur-
face plasmon resonance is strongly dependent on their particle size and morphology.
Therefore, gold- and silver-based noble metals are extensively studied with various
semiconductor photocatalysts towards improving the overall photocatalytic perfor-
mance [20]. For example, the Ag/Bi3TaO7 plasmonic photocatalyst showed signifi-
cantly higher photocatalytic efficiency up to 85.42% tetracycline removal in 120 min
under visible light irradiation, which was ∼1.5-fold higher as compared to the pris-
tine Bi3TaO7. The improved photocatalytic efficiency was endorsed to the surface
plasmon resonance effect of metallic silver (Ag) nanoparticles. Meanwhile, the
potential incorporation of Ag nanoparticles into Bi3TaO7 surface established the
inner electromagnetic field, which boosted the photogenerated charge carrier sepa-
ration (Fig. 4.3) [25].
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 91

Fig. 4.3  Photocatalytic mechanism of Ag/Bi3TaO7 plasmonic photocatalyst for degradation of


tetracycline [25]

The Ag/Ag2MoO4 composite showed significantly improved photocatalytic


activity under visible light due to the SPR effect of Ag nanoparticles that anchored
on the surface of Ag2MoO4. Also, Ag/Ag2MoO4 composite revealed a greatly
improved photocatalytic activity as compared to pure Ag2MoO4, which presented
around 99.5% of ciprofloxacin removal in 60  min and excellent stability for five
successive cycles [26]. Further, the silver-incorporated Bi3O4Cl (Ag/Bi3O4Cl) com-
posite presented an enhanced photocatalytic activity on tetracycline removal, where
it found to degrade 94.2% of tetracycline in 120 min under visible light. The boosted
photocatalytic activity of Ag/Bi3O4Cl composite was attributed to the SPR absor-
bance of silver nanoparticles [27]. Likewise, Au nanoparticles (~10 nm) were uni-
formly dispersed on ZnO microstructure, and this system showed potential
photocatalytic activity on tetracycline degradation, where around 85.5% of tetracy-
cline was degraded in 120 min under visible light irradiation. Moreover, the control-
lable size and loading of Au nanoparticles on ZnO surface was found to tune the
plasmonic and synergetic effect between Au and ZnO as well as their overall photo-
catalytic efficiency [28].
Further, a single atom silver (Ag)-loaded ultrathin g-C3N4 hybrid (AgTCM/
UCN) was fabricated via a facile co-polymerization method and studied for the
degradation of sulfamethazine in the presence of peroxymonosulfate (PMS) under
visible light. The photocatalytic results displayed that the AgTCM/UCN/PMS sys-
tem exhibits higher photocatalytic activity than that of the AgTCM/UCN, UCN/
PMS, and g-C3N4/PMS. The observed enhanced photocatalytic activity was ascribed
to the combined effects of SPR to Ag, efficient charge separation of PMS, and high
surface area of UCN. Then, the radical test revealed that the reactive species SO4˙−,
O2˙−, and h+ were found to play significant roles on the photocatalytic degradation
of sulfamethazine [29]. Further, BiVO4/Ag/Cu2O nanocomposite displayed
enhanced photocatalytic performance towards tetracycline removal, where 91.22%
of removal efficiency was observed within 60 min under visible light irradiation.
Besides, this nanocomposite also showed enhanced mineralization ability of around
55.32% even after 60  min. The excellent photocatalytic degradation and
92 R. Karuppannan et al.

mineralization was established by the introduction of Cu2O and plasmonic Ag,


where it improved the visible light absorption and SPR effect of metallic Ag. It col-
lectively led to a dual charge transfer pathway in this composite system [30].
Similarly, the Au and CuS nanoparticle-modified TiO2 nanobelt (Au–CuS–TiO2
NBs) ternary composite was successfully employed for oxytetracycline degrada-
tion, where it demonstrated a superior photocatalytic performance of 96% degrada-
tion in 60 min under simulated solar irradiation. The superior activity was attributed
to the plasmonic effect of Au and CuS nanoparticles (induced by Cu defects), where
it collectively enhanced the visible light utilization, and the synergetic effect among
the Au–CuS–TiO2 ternary composite remarkably hindered the photogenerated
electron-­hole charge recombination [31]. Further, the plasmonic gold nanoparticle-­
loaded porous graphitic nitride/graphene-layered (Au/pg–C3N4/GR) ternary plas-
monic photocatalyst was explored for ciprofloxacin degradation under visible light
irradiation. The ternary plasmonic composite was found to show 4.34 and 3.05
times higher degradation performance than that of porous g-C3N4. Moreover, the
tunable surface plasmon resonance (SPR) effect of Au and electron-acceptor nature
of graphene potentially improved the visible light utilization and also facilitated the
higher photoinduced charge carrier separation [32].

4.5.2  Semiconductor–Semiconductor Composites

The semiconductor–semiconductor-based heterojunction construction offers highly


distinctive charge collection and separation; thereby it drastically improves the pho-
tocatalytic efficiency of the composite. Besides, the designing of semiconductor–
semiconductor composite provides several benefits such as potential charge
separation and superior charge transfer, prolonged charge carrier lifetime, separate
active sites, and extended light absorbance. The semiconductor-semiconductor het-
erostructure composite is typically constructed by assembling two different semi-
conductors and used to be well connected via effective interfacial contacts. Further,
the semiconductor–semiconductor heterojunction composite is mostly established
with two different potential architectures, known as p-n heterojunction and Z-scheme
heterojunction towards improving the photocatalytic performance of the composite
systems [33].

p-n Heterojunction Composites

The p-n heterojunction photocatalyst is constructed by coupling the p-type and


n-type semiconductors (SCs). In the p-n system, semiconductors with two different
charge interfaces create efficient integral potential with easy charge separation,
which delivers fast and efficient electron-hole pair separation and transfer. In p-n
junction, the photogenerated electron transfer from SC with more negative conduc-
tion band potential into semiconductor with lesser negative conduction band and the
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 93

photogenerated holes migrate from SC with a lower valence band to semiconductor


with higher valence band. Therefore, the construction of p-n junction effectively
promotes the charge separation and inhibits the charge recombination thereby
improving the photocatalytic efficiency [34]. For example, the BiOI/La2Ti2O7 p-n
heterojunction composites presented much higher photocatalytic efficiency than
pure La2Ti2O7 towards the ciprofloxacin degradation under visible light irradiation.
The BiOI/La2Ti2O7 composite was found to show 274 times higher ciprofloxacin
removal rate as compared to pure La2Ti2O7. The outstanding photocatalytic effi-
ciency was attributed to the formation of p-n heterojunctions between BiOI and
La2Ti2O7, where it improved the photogenerated charge carrier separation and trans-
fer efficiency. In addition, the narrow bandgap of BiOI improved the photo-response
of La2Ti2O7 in the visible light region [35]. Further, the marigold flowerlike hierar-
chical architecture of CoO@MnCo2O4 p-n heterojunction composite was fabricated
by anchoring CoO nanoparticles on the surface of MnCo2O4. The as-developed
CoO@MnCo2O4 p-n heterojunction established efficient photoinduced charge car-
rier separation process by the formation of p-n heterojunction complex with strong
inner electric field. Moreover, the unique three-dimensional hierarchical micro-­
flower structure delivered high specific surface area and sufficient active sites for the
photocatalytic reaction. Therefore, CoO@MnCo2O4 hybrid composite was found to
show an efficient photocatalytic oxidation of tetracycline and excellent photocata-
lytic reduction of hexavalent chromium under visible light irradiation. Further, this
photocatalyst system was also inspected for the treatment a mixture of pollutants
and found higher photocatalytic efficiencies (Fig. 4.4) [36].
The Co3O4–C3N4 p-n nano-heterojunctions with appropriate band bending were
fabricated by dispersing the ultrafine Co3O4 nanoparticles onto the g-C3N4 sheets.
Further, this well-constructed p-n nano-heterojunction was found to significantly
reduce the photoexcited charge carrier’s recombination and resulted into higher
photocatalytic degradation of mixture of tetracycline and methylene blue pollutants
under sunlight irradiation. Besides, the Co3O4–C3N4 p-n nanocomposite was found
to have tenfold higher surface area as compared to the pristine C3N4 and also showed

Fig. 4.4  The proposed photocatalytic charge transfer mechanism of CoO@MnCo2O4p-n hetero-
junction [36]
94 R. Karuppannan et al.

extended visible light absorption due to the presence of Co3O4 nanoparticles [37].
Similarly, a stable p-n heterojunction CuS/BiVO4 photocatalyst through in suit
growth of CuS nanoparticles over the surface of BiVO4 was developed, and it
showed an enhanced visible light absorption and effectively promoted the photoin-
duced charge carrier’s separation. The CuS/BiVO4 heterostructure also offered high
surface area and large number surface-active sites for the effective photocatalytic
process. Accordingly, the CuS/BiVO4 composites showed 86.7% ciprofloxacin deg-
radation under visible light. The degradation efficiency was 2.59 and 16.54 times
higher than that of pristine BiVO4 and CuS, respectively. Also, the holes (h+) were
identified as main active species in ciprofloxacin degradation and the possible deg-
radation pathway was also proposed [38]. Similarly, an in situ loading of Ag2O
particles on CeO2 spindles’ surface was developed to construct the Ag2O/CeO2 p-n
heterojunction composites, and their photocatalytic activity was investigated on the
degradation of enrofloxacin under visible light irradiation. The p-n heterojunction
formation offered higher photoinduced charge carrier separation efficiency in Ag2O/
CeO2 composites; therefore the improved photocatalytic degradation efficiency of
87.11% in 120 min and high mineralization efficiency of 66.82% in 160 min for
antibiotic enrofloxacin were observed [39]. Further, the CdS QDs decorated
p-CaFe2O4@n-ZnFe2O4 ternary heterojunction hybrid composite was developed as
a potential photocatalyst towards the degradation of norfloxacin under solar light
irradiation. This hybrid ternary composite showed 83% degradation of norfloxacin
in 90 min, which was 1.28 times higher than that of CaFe2O4@ZnFe2O4. The p-n
modified bandgap alignment in p-CaFe2O4@n-ZnFe2O4 offered the heterojunction
charge transfer and isoenergetic electron transfer from CdS QDs to CaFe2O4, and
therefore the higher photoinduced charge generation and transfer was established in
CdS QDs/CaFe2O4@ZnFe2O4 ternary nanocomposite [40]. Similarly, a novel plas-
monic p-n heterojunction photocatalyst Ag/Ag2O/PbBiO2Br was potentially devel-
oped and studied for the tetracycline degradation. The synergistic effect between
plasmonic metallic Ag and Ag2O/PbBiO2Br p-n heterojunction greatly enhanced the
photo-utilization and tremendously accelerated the charge separation. Therefore,
the Ag/Ag2O/PbBiO2Br heterojunction plasmonic composite demonstrated around
84.4% (90 min) and 50.9% (240 min) of tetracycline degradation under visible light
irradiation and NIR light irradiation, respectively. Meanwhile, the O2˙−, h+, and •OH
species were found to be the main active species under visible light, while O2˙− and
h+ under NIR light reaction. In addition, the tetracycline degradation pathway was
also proposed as shown in Fig. 4.5 [41].
Interestingly, the nitrogen-doped graphene quantum dots (NGQDs)-BiOI/
MnNb2O6 p-n junction composite was developed and employed for the degradation
of multiple antibiotics such as tetracycline, oxytetracycline, ciprofloxacin, and dox-
ycycline. Accordingly, the developed composite was found to show higher rate of
photocatalytic activity as compared to the pristine MnNb2O6 and BiOI. The higher
photocatalytic performance was attributed to the combined effect of p-n junction
BiOI/MnNb2O6 and NGQDs, which promoted the charge carrier mobility and
thereby greatly improved the electron-hole pair charge separation as well as reduced
the charge carrier recombination. Importantly, the Bi/Mn ratio and the
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 95

Fig. 4.5  The proposed degradation pathway of tetracycline in the presence of plasmonic Ag/
Ag2O/PbBiO2Br p-n heterostructure under visible light irradiation [41]

concentration of NGQDs were found to play important roles in the photocatalytic


efficiency of this system [42].

Z-Scheme Heterojunction Composites

The Z-scheme heterojunction materials possess great consideration in photocata-


lytic applications due to their stronger reduction and oxidation capacity, which
essentially offered superior photocatalytic performances. The Z-scheme photocata-
lysts own several advantages such as separate reductive and oxidative active sites,
higher charge carrier separation and charge transfer, and wide range of photoabsor-
bance. As compared to p-n heterojunction composites, the Z-scheme composites
possess the same band structure but with reverse charge transfer mechanism. The
oxidation reaction occurs in semiconductor at lower valence band and reduction
reaction occurs at higher conduction band. As a result, the Z-scheme photocatalysts
possess separate reduction and oxidation sites as well as lesser charge recombina-
tion towards photocatalytic reactions [43, 44].

All-Solid-State Z-Scheme Heterojunction Composites

The all-solid-state Z-scheme or mediator-containing Z-scheme photocatalysts can


be designed by integrating two different semiconductors with a solid electron medi-
ator at the interface of the semiconductors. The photogenerated electron transfer
96 R. Karuppannan et al.

occurs from lower conduction band of the semiconductor to solid-state mediator,


and then the electrons are effectively trapped by higher valence band of the semi-
conductor. As a result, it leads to the continuous redox reaction at solid-state media-
tor. Therefore, the higher rate of charge carrier separation and transfer is succeeded
by the solid-state mediator, which is usually the noble metals such as Au and Ag and
graphene materials like reduced graphene oxide (RGO) [43]. For example, the
mediator-containing Z-scheme BiOCl–Au–CdS heterostructure was developed
through the selective photo-deposition of Au nanoparticles on BiOCl nanosheets,
and then the CdS nanoparticles were in situ anchored on Au NPs via strong S-Au
(sulfur-gold) interaction. The obtained results concluded that the BiOCl–Au–CdS
Z-scheme structure offered lesser photogenerated electron-hole recombination with
higher redox capacity. Therefore, the as-developed Z-scheme BiOCl-Au-CdS het-
erostructure established significantly higher photocatalytic sulfadiazine degradation
as compared to bare BiOCl, BiOCl-Au, and BiOCl-CdS materials under solar light
[45]. Similarly, the in situ formations of Ag@AgI nanoparticles on the surface of
defective BiOI nanosheets were achieved to construct the iodine vacancy-rich BiOI/
Ag@AgI (Ag@AgI/VI-BOI) Z-scheme heterojunction photocatalyst. This as-­
developed Z-scheme Ag@AgI/VI-BOI heterojunction photocatalyst was found to
be an efficient visible light-driven photocatalytic system, which degraded around
86.4% of tetracycline in 60 min. The combined effects of charge transfer mecha-
nism in silver mediator containing Z-scheme and the iodine vacancies (which pro-
moted the effective electron-hole pair charge separation and the suppression of
photoinduced charge recombinations) were attributed to their enhanced photocata-
lytic tetracycline degradations. Further, it was observed that the in situ synthesis
offered the effective interfacial contact between VI-BOI and Ag@AgI as compared
to the physically mixed synthesis of Ag@AgI/VI-BOI nanocomposite system [46].
Recently, the plasmonic metal Ag-bridged 2D/2D Bi5FeTi3O15/g-C3N4 Z-scheme
heterojunction photocatalyst with powerful interfacial charge transfer was estab-
lished through ultrasound process. The as-developed 2D/2D heterostructure com-
posite exhibited a notable improved photocatalytic activity on tetracycline
degradation under visible light (~70% in 60  min) and simulated solar irradiation
(~85% in 20  min). It was explained that the incorporation of Ag nanoparticles
between the interlayers of Bi5FeTi3O15/ultrathin g-C3N4 nanosheets established the
high-speed charge transfer channel and accelerated the charge transfer in the con-
structed 2D/2D heterostructure (Fig. 4.6) [47].
The oxygen-deficient Ag-deposited WO3−X nanocapsules were coupled with Zn/
Cr LDH to develop the mediator-containing Z-scheme WO3−X/Ag/ZnCr LDH het-
erostructure photocatalyst. The strong internal electric field-induced Z-scheme
charge transfer mechanism was responsible for the enhanced photocatalytic perfor-
mance of heterostructure and showed 92% tetracycline removal in 90  min under
visible light. Besides, the oxygen vacancies in WO3−X contributed to increase the
photoinduced charge carrier’s lifetime and enhanced antibiotic adsorption on cata-
lyst surface. Furthermore, the Ag NPs acted as electron conduction bridge, where it
played an important role on the charge transfer mechanism of the Z-scheme photo-
catalysts [48]. Further, the graphene-bridged Ag3PO4/Ag/BiVO4 Z-scheme
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 97

Fig. 4.6  The Z-scheme superfast interfacial charge transfer of 2D/2D Bi5FeTi3O15/g-C3N4 hetero-
structure [47]

heterojunction photocatalyst with excellent visible light performance was con-


structed by an in situ deposition and photoreduction method. The obtained hetero-
junction photocatalyst was employed for degradation of tetracycline and displayed
superior photodegradation efficiency of 94.96% tetracycline removal in 60 min. The
enhanced catalytic performance was attributed by Ag mediator-induced Z-scheme
mechanism, which offered better charge separation and desirable visible absorption.
The RGO contributed on the improvement of photostability and recyclability of
Ag3PO4/Ag/BiVO4 heterojunction composite [49]. Similarly, the introduction of
Ag/PEDOT into ZnFe2O4 surface (to construct Z-scheme structure) was found to
enhance the photocatalytic activity, where the existence of imprinted cavity on the
surface significantly improved the selectivity of the tetracycline. As a result, the
photocatalytic activity of Z-scheme ZnFe2O4/Ag/PEDOT was demonstrated on the
selective degradation of antibiotic tetracycline under the simulated sunlight, which
showed around 1.32 times higher efficiency than that of the non-imprinted ZnFe2O4/
Ag/PEDOT system. Moreover, the as-prepared imprinted ZnFe2O4/Ag/PEDOT
materials possessed good hollow spherical structure, improved light response abil-
ity, magnetic separation, and stability [50].
On the other hand, the construction of natural photosynthesis-inspired Z-scheme
tandem photocatalyst provided a promising strategy for the efficient electron-hole
separation and charge transfer, where it also provided a unique prospect to achieve
higher photocatalytic performance for antibiotic degradation. Furthermore, it is
believed that an ideal electron mediator is critical to promote the electron transfer in
Z-scheme systems. For example, the well-defined CeVO4/3D RGO aerogel/BiVO4
ternary composite was fabricated and delivered a greater interfacial contact and
higher tetracycline photodegradation. Furthermore, the tetracycline degradation
efficiency of CeVO4/3D RGO /BiVO4 ternary composite was found 3.7 times higher
than that of BiVO4 under visible light irradiation. The higher photocatalytic activity
was established by the 3D RGO aerogel, which acted as an efficient electron media-
tor bridge between the two vanadate structures and offered more efficient tandem
charge transfer channels (Fig.  4.7). In addition, the specific 3D heterostructure
offered superior benefits to the adsorption and photodegradation of tetracycline [51].
98 R. Karuppannan et al.

Fig. 4.7  The proposed


Z-scheme charge transfer
mechanism of BiVO4/
RGO/CeVO4 ternary
composite [51]

Direct Z-Scheme Heterojunction Composites

The direct Z-scheme or mediator-free Z-scheme heterojunctions are designed by


combining two different semiconductors without the incorporation of any mediator
materials. Such system contains two types of semiconductors, where one semicon-
ductor is responsible for strong oxidation and the other is for strong reduction reac-
tions. The semiconductor with lower valence band potential establishes strong
oxidation ability and performs the oxidation reactions. The semiconductor with
higher conduction band possesses strong reduction ability and performs reduction
reactions. The direct Z-scheme systems are relatively more advantageous as com-
pared to the mediator-containing Z-scheme system as they provide the limited back-
ward reaction, higher stability, and fast and simple charge transfer [44]. For example,
the redox mediator-free direct Z-scheme AgI/WO3 nanocomposite was successfully
constructed via a facile precipitation method. The formation of direct Z-scheme
heterojunction was found to facilitate the charge migration from WO3 to
AgI. Therefore, AgI/WO3 Z-scheme nanocomposite exhibited an enhanced photo-
catalytic activity towards the degradation of tetracycline, which is over 4.3 and 25
times higher than that of pure AgI and WO3, respectively [52]. Similarly, the WO3-­
g-­C3N4 mediator-free Z-scheme composite showed 91.7% removal of sulfamethox-
azole under visible light. The enhanced catalytic performance was assigned to the
well-established band structure alignment between g-C3N4 and WO3 and the
Z-scheme charge transfer mechanism in WO3-g-C3N4 heterojunction composite. It
was observed that the electrons in the conduction band of g-C3N4 showed high
reduction capacity, and the holes in the valence band of WO3 showed higher oxidi-
zation capacity [53]. The Z-scheme mes-Sn3O4/g-C3N4 heterostructure composite
was found to degrade 72.2% of tetracycline and mineralize 61.2% of tetracycline
under visible light irradiation. This developed Z-scheme heterostructure favored for
the enhanced charge separation and effectively reduced the charge carrier recombi-
nation between Sn3O4 and g-C3N4; thereby it greatly enhanced the photocatalytic
activity. In addition, the mesoporous structure and enhanced specific surface area of
Z-scheme Sn3O4/g-C3N4 heterostructure provided abundant active sites to absorb
more reactant molecules and thereby improved the photocatalytic activity [54].
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 99

Similarly, the InVO4 hollow spheres were uniformly loaded on the surface of g-C3N4
nanorod to construct g-C3N4/InVO4 heterojunction Z-scheme composite with an
intimate interface. The unique structural properties of g-C3N4/InVO4 and Z-scheme
mechanism promoted excellent visible light absorption and greatly reduced the pho-
togenerated electron-hole pair recombination and prolonged the charge carriers’
lifetime. These improved properties collectively enhanced the photocatalytic activ-
ity towards tetracycline degradation as compared to the bare g-C3N4and InVO4 [55].
Similarly, an efficient direct Z-scheme based on CuInS2/Bi2WO6 heterojunction
with intimate interface contact was constructed via the direct growth of Bi2WO6 on
CuInS2 microspheres. The efficient intimate interface contact offered direct charge
transfer pathways in the CuInS2/Bi2WO6 Z-scheme heterojunction. Such arrange-
ment remarkably promoted the photogenerated electron-hole separation and led to
the higher photocatalytic performance (Fig.  4.8). Accordingly, the developed
Z-scheme CuInS2/Bi2WO6 heterojunctions showed a higher tetracycline removal
efficiency (90.5%) as compared to pure CuInS2 and Bi2WO6 under visible light
irradiation [56].
Similarly, LaFeO3 nanoparticles were well dispersed on SnS2 nanosheets to con-
struct a binary Z-scheme LaFeO3/SnS2 heterojunction with strong interfacial con-
tacts. The Z-scheme LaFeO3/SnS2 heterojunction showed extended visible light
absorption and facilitated the higher photoinduced charge separation and transfer
efficiency. Accordingly, the constructed hybrid LaFeO3/SnS2 photocatalyst showed
3.5 and 2.2 times higher tetracycline degradation than that of the bare LaFeO3 and
SnS2 under visible light [57]. Further, the Z-scheme core-shell-type Fe2O3/CuBi2O4
heterostructure was constructed, and it showed higher tetracycline degradation effi-
ciency as compared to the pure Fe2O3 and CuBi2O4. Furthermore, the Fe2O3/CuBi2O4

Fig. 4.8  The interfacial electron transfer process and possible photocatalytic mechanism of
Z-scheme CuInS2/Bi2WO6 heterojunction [56]
100 R. Karuppannan et al.

composite showed highest degradation rate of nearly 80% in 120 min with outstand-
ing stability. This higher photocatalytic performance and stability was mainly
ascribed to their core-shell-like structure of Z-scheme Fe2O3/CuBi2O4 composite
[58]. Similar to this, the construction of 0D/2D mediator-free Z-scheme photocata-
lytic system was constructed, and it showed drastically improved photoinduced
charge carrier separation as well as the enhanced photocatalytic performance. The
uniform dispersion of zero-dimensional Bi3TaO7 quantum dots on two-dimensional
g-C3N4 nanosheets was performed to fabricate this visible light-responsive
Bi3TaO7QDs/g-C3N4NSs 0D/2D Z-scheme heterostructure and explored for the cip-
rofloxacin degradation. It showed 91% degradation in 120 min, which was around
four and 12.2 times higher photocatalytic performance than the bare Bi3TaO7 and
g-C3N4, respectively. The observed enhanced photocatalytic activity was attributed
to the strong interaction between Bi3TaO7 QDs and g-C3N4 NSs, where it estab-
lished an efficient 0D/2D Z-scheme charge separation and transfer [59]. Similarly,
a 2D/2D Z-scheme heterojunction WO3/K+Ca2Nb3O10− photocatalyst was con-
structed using hydrothermal method and explored for the photocatalytic degrada-
tion of tetracycline hydrochloride under simulated sunlight. The increased specific
surface area and strong heterojunction interfacial contacts were attributed to their
enhanced photoinduced charge carrier separation and charge transfer. Accordingly,
the WO3/K+Ca2Nb3O10− heterojunction composite showed 5.1-fold and twofold
higher efficiency than that of pristine WO3 and K+Ca2Nb3O10− nanosheets, respec-
tively [60]. Further, a 2D/3D g-C3N4@ZnO heterostructure Z-scheme catalyst was
developed and studied for the degradation of cephalexin, where it degraded around
98.90% and 72.8% of cephalexin in 60 min under simulated solar light. The devel-
oped Z-scheme g-C3N4@ZnO heterostructure achieved 5.4 and 8.1 times more deg-
radation rate than that of the bare g-C3N4 and ZnO, respectively. The observed
higher photocatalytic activity was mainly ascribed to the strong oxidation capacity
of the formed active species generated by the developed Z-scheme heterostructure.
Also, the consecutive degradation experiments revealed their excellent stability and
recyclability of 2D/3D g-C3N4@ZnO heterostructure system. Moreover, the by-­
products of cephalexin degradation were also identified and the possible degrada-
tion pathway of cephalexin was also proposed (Fig. 4.9) [61].

Other Heterojunction Composites

The designing of semiconductor-semiconductor-based heterojunctions between the


two unequal band structure semiconductors is one of the facile approaches to
improve the photocatalytic efficiency. The heterojunction photocatalyst offers effec-
tive electron-hole pair separation and charge transfer without any restriction.
Actually, the heterojunction charge transfer mechanism is distinct than the conven-
tional p-n junction and Z-scheme and heterostructure photocatalytic system [62].
For example, a magnetically recoverable BiOBr/Fe3O4 heterostructure showed out-
standing photocatalytic performance on the norfloxacin degradation, and it degraded
more than 90% norfloxacin within 30  min under visible light irradiation. The
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 101

Fig. 4.9  The possible degradation pathway and intermediates of cephalexin [61]

formation of this heterojunction led to an efficient electron-hole separation, which


minimized the photoexcited electron-hole pair recombination; thereby it showed the
improved photocatalytic efficiency. Moreover, the observed enhanced photocata-
lytic activity was strongly dependent on the band structure of the heterostructure
photocatalytic materials. Also, the radical-trapping experiments revealed that the
photodegradation of norfloxacin was dominated by the superoxide radicals and
hydroxyl radicals rather than direct holes [63]. Further, the novel 3D microsphere-­
like In2S3/InVO4 heterojunction showed around 2.26 and 11.71 times higher tetracy-
cline degradation efficiency than that of the individual In2S3 and InVO4, respectively.
Owing to the close interconnection between In2S3 and InVO4, their interface has
significantly improved the charge separation and inhibited the recombination of
photogenerated electron-hole pairs. Also, the photosensitization of In2S3on InVO4
expanded the overall optical absorption of the In2S3/InVO4 composite and improved
their photostability as well [64]. The ultrafine 0D TiO2 nanoparticles were well dis-
persed on 2D SnNb2O6 nanosheet to fabricate TiO2/SnNb2O6 heterojunction photo-
catalyst. The as-fabricated TiO2/SnNb2O6 photocatalyst was found to have a suitable
wide and narrow bandgap energy and demonstrated higher photocatalytic activity
for the degradation of tetracycline under visible light. The formation of significant
interfacial contact and identical band energy between heterojunctions improved the
charge transfer efficiency and suppressed electron-hole recombination of SnNb2O6
under visible light (Fig. 4.10) [65].
Similarly, the visible light-driven 2D–2D heterojunction was developed by com-
bining the 2D g-C3N4 and 2D ultrathin K+Ca2Nb3O10− perovskite nanosheets and
102 R. Karuppannan et al.

Fig. 4.10  The proposed photocatalytic charge transfer mechanism of TiO2/SnNb2O6 heterojunc-
tion under visible light irradiation [65]

studied for tetracycline degradation. The formation of 2D-2D heterojunctions with


strong interfacial contact enhanced the photoinduced electron transfer, and there-
fore, the as-developed g-C3N4/K+Ca2Nb3O10− heterojunction showed 81% degrada-
tion of tetracycline in 90 min, where it was about 6.6 and 1.8 times higher than that
of K+Ca2Nb3O10− and g-C3N4, respectively [66]. The carbon nitride-modified hierar-
chical yolk-shell-like titanium dioxide sphere (g-C3N4/TiO2) heterojunction photo-
catalysts showed remarkable visible photocatalytic activity towards ciprofloxacin
degradation. The experimental results showed that the heterojunction composite
g-C3N4/TiO2 was found to have greater photocatalytic degradation efficiency
(97.3%) as compared to the pure TiO2 hollow sphere (53.5%) and g-C3N4 (19.4%).
This observed efficiency was attributed to strong interfacial contact between TiO2
and g-C3N4, which favored for the enhanced charge carrier separation. Furthermore,
the unique multiple light reflection property of TiO2 hollow microspheres within the
interior cavities increased their light absorption capacity and thereby improved the
overall photocatalytic activity [67]. Interestingly, the rational designing of hollow
porous Co2SnO4-SnO2/graphite carbon (Co2SnO4–SnO2/GC) nanotube multi-­
heterojunction demonstrated a remarkable photocatalytic activity on the degrada-
tion of chlortetracycline (~83% in 80 min) and tetracycline (~80% in 80 min) under
visible light irradiation. The synergistic effects between multi-junctions notably
improved their electron-hole pair charge generation and transfer; thereby it signifi-
cantly inhibited the charge recombination rate and improved the catalytic activity.
In addition, the graphite carbon maintained the overall photoactivity and stability of
Co2SnO4-SnO2/GC heterojunction composites [68].

4.5.3  Semiconductor–Carbon Composites

The carbonaceous materials such as graphene (GO/RGO), carbon quantum dots


(CQDs), and carbon sphere are widely explored in photocatalysis due to their poten-
tial properties towards efficient charge separation, higher charge transfer, extended
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 103

photoabsorbance, high specific surface area, and good stability. Therefore,


semiconductor-­carbon-based composites have also been explored to improve the
overall photocatalytic performances of the systems [69, 70].

Semiconductor–Graphene Composites

The graphene materials have received huge attraction in hybrid photocatalytic mate-
rial designing because of their unique two-dimensional hybridized carbon structure,
which essentially offers higher specific surface area, fast charge transfer, and greater
chemical stability. Also, their unique structural properties offer efficient photoin-
duced charge separation and transport properties, thereby establishing the improved
photocatalytic performances [71]. Therefore, graphene materials such as graphene
oxide (GO) and reduced graphene oxide (rGO)-based composites have received
much attention in photocatalytic antibiotics degradation. For instance, the graphene
oxide-modified lanthanum vanadate (GO/LaVO4) nanocomposite was potentially
used to remove the tetracycline and naproxen under visible light irradiation. This
developed GO/LaVO4 nanocomposite revealed higher photocatalytic degradation
efficiency, which was about 3.46 (tetracycline) and 2.29 (naproxen) times higher
than that of pure LaVO4. The greater photocatalytic performance was attributed to
the accelerated photoinduced charge separation and transfer by GO and the improved
optical absorption by LaVO4 [72]. Similarly, Ce(MoO4)2 nanocube/graphene oxide
(CeM/GO) composite was developed for the efficient photocatalytic degradation and
electrochemical reduction of neurotoxicity antibiotic chloramphenicol molecules.
The CeM/GO nanocomposite showed an excellent photocatalytic performance
under visible light with 99% degradation of chloramphenicol within 50 min. The
observed outstanding performance was attributed to the GO-induced efficient charge
separation and migration in the CeM/GO composite [73]. Similarly, it was found
that the reduced graphene oxide (RGO) effectively transports the electrons and
reduces the electron-hole recombination rate in RGO–CdS/ZnS composite.
Accordingly, the synthesized RGO–CdS/ZnS composites showed higher photocata-
lytic performance on the degradation of tetracycline as compared to the CdS, ZnS,
and CdS/ZnS materials. Also, RGO acted as a good catalytic supporter, which pre-
vented the CdS/ZnS from photo-corrosion and thus provided the improved stability
and catalytic efficiency [74]. Further, the sandwich-like TiO2–rGO composite was
fabricated via Pickering emulsion approach, where the TiO2 nanoparticles were
closely and densely packed on rGO sheets (Fig. 4.11). This unique structure signifi-
cantly enhanced the light absorption, accelerated the charge separation, and improved
the adsorption capacity of antibiotic, thereby increasing the photocatalytic removal
of tetracycline under visible light around, which was around 94% in 40 min [75].
Further, the nitrogen-doped graphene quantum dots (N-GQDs) and plasmonic
silver (Ag NPs) co-decorated ultrathin graphitic carbon nitride (g-C3N4) nanosheet
composites were developed and demonstrated towards the degradation of tetracy-
cline. The Ag/N-GQDs/g-C3N4 composite reached 92.8% of tetracycline removal in
60  min, where the observed higher photocatalytic activity was attributed to the
104 R. Karuppannan et al.

Fig. 4.11  Fabrication of the TiO2–GO composite by Pickering emulsion method [75]

synergistic effect between the g-C3N4, N-GQDs, and Ag NPs. Moreover, it was
found that the N-GQDs acted as an efficient charge transporter and electron sink,
and thereby it improved the photoinduced electron migration and inhibited the
charge recombination. Also, Ag NPs were found to promote the light absorption and
transfer ability, leading to the generation of more photoinduced charge carriers
towards the effective degradation of tetracycline [76].

Semiconductor-Carbon Quantum Dot Composites

The carbon quantum dots (CQDs) are a unique material, where they act as an elec-
tron acceptor as well as electron donor, and thus the semiconductor-CQDs estab-
lished a great consideration in photocatalytic applications. Furthermore, the CQDs
can also act as photosensitizers, through which they offer the extended photoabsor-
bance in semiconductor in the visible light region [77]. For example, the photocata-
lytic performance of CQD-modified Bi2MoO6 was explored on the degradation of
ciprofloxacin and tetracycline hydrochloride under visible light. The well distribu-
tion of CQDs (~7 nm) on Bi2MoO6 nanosheet surface enhanced the photocatalytic
activity and found to have 5.7 times higher ciprofloxacin degradation efficiency than
that of pure BiMoO6. The CQDs played a major role on the observed superior pho-
tocatalytic activity of CQD-BiMoO6 composite. In this system, the CQDs essen-
tially acted as a photocenter for light absorbance as well as the charge centers
towards improving the charge separation and reducing the charge recombination and
thereby enhanced the antibiotic degradation [78]. Similarly, the CQDs/BiOI com-
posite showed excellent photocatalytic activities than that of the pure BiOI towards
the degradation of tetracycline under visible light. The CQDs offered enhanced vis-
ible light absorption, increased specific surface area, and improved photogenerated
electron-hole charge separation efficiency. In addition, the introduction of CQDs
induced the formation of ultra-small BiOI nanosheets with assembled hollow micro-
sphere structure. This unique structural feature was also found to be one of the
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 105

essential reasons that offered the excellent photocatalytic activity [79]. Likewise,
12 nm-sized carbon dots (C-dots) were successfully dispersed over the TiO2 surface
to fabricate TiO2/C-dot composite and applied for antibiotic levofloxacin removal
under solar light irradiation. The TiO2/C-dot composite showed almost complete
degradation of levofloxacin within 90 min, which was found to be superior as com-
pared to the bare and commercial TiO2 catalysts. Because of the presence of C-dots
over TiO2, it enhanced the photo-absorption and offered efficient photoinduced
charge separation in TiO2/C-dot composite. In addition, the presence of the C-dots
in the system provided more active sites for surface adsorption of levofloxacin mol-
ecules as well as restricted the recombination of charge carriers. Furthermore, the
degradation products and intermediates of levofloxacin were also identified, and the
possible degradation pathway was proposed in detail [77]. Interestingly, the multidi-
mensional carbon quantum dots@3D daisy-like In2S3/single-wall carbon nanotubes
(CQDs@In2S3/SWNTs) semiconductor-carbon heterostructure composite demon-
strated a fast charge transfer and improved visible photocatalytic activity for the
degradation of antibiotics ciprofloxacin, tetracycline, and levofloxacin. Thereby, the
CQDs offered superior charge separation and transfer abilities and improved the
photocatalytic activity. In addition, the plasma effect of CQDs significantly influ-
enced the photocatalytic performance. Furthermore, the 3D multidimensional hier-
archical structure led to the enhancement of light utilization properties of CQDs@
In2S3/SWNTs composite. Also, the radicals such as O2•− and h+ were identified as
main active species in the photocatalytic degradation. Accordingly, the detailed deg-
radation mechanism of antibiotics was also proposed as shown in Fig. 4.12 [80].

Semiconductor–Carbon Sphere Composites

Recently, the carbon sphere-based composites have received significant importance


in photocatalytic material designing because of their low cost and simple synthesis
procedures. Importantly, these carbon spheres can also act as efficient supporting/
template materials for the encapsulation of semiconductor particles; thereby they
enhance the surface properties of the semiconductor composite photocatalyst [81].
For example, the modified carbon spheres/g-C3N4 composites showed efficient pho-
tocatalytic activity as compared to the unmodified carbon spheres/g-C3N4 compos-
ite and g-C3N4, which showed around 98% of tylosin degradation within 45 min
under simulated solar light. It was found that the higher surface of carbon sphere
and formation of potential heterojunction with closer interface contact between car-
bon spheres and g-C3N4 promoted the electron-hole pair separation and boosted up
the visible light absorption in carbon sphere/g-C3N4 composites. In addition, the
CTAB-modified carbon spheres found to have a higher degree of carbonation and
effectively transfer the electrons at faster rates as compared to the unmodified car-
bon spheres. Notably, in this study, the carbon spheres were synthesized by hydro-
thermal reaction using different carbon precursors such as glucose, β-cyclodextrin,
and sucrose with and without adding the surfactant CTAB as shown in Fig. 4.13 [82].
Further, a direct dual semiconductor photocatalyst based on the nitrogen-
doped  hollow mesoporous carbon sphere-modified g-C3N4/Bi2O3 (g-C3N4/
106 R. Karuppannan et al.

Fig. 4.12  The photocatalytic degradation mechanism of CQDs@In2S3/SWNTs composites under


visible light irradiation [80]

Fig. 4.13  The development of modified carbon sphere/graphitic nitride (T-CS/g-C3N4) compos-
ites [82]

Bi2O3@N-­HMCs) was demonstrated for the degradation of tetracycline hydrochlo-


ride and ciprofloxacin hydrochloride under visible light irradiation. The g-C3N4/
Bi2O3@N-HMCs composite exhibited superior photocatalytic efficiency as com-
pared to the single-­phase g-C3N4, Bi2O3 and binary composites (g-C3N4/N-HMCs,
Bi2O3/N-HMCs, and g-C3N4/Bi2O3). The observed enhanced photocatalytic effi-
ciency of g-C3N4/Bi2O3@N-HMCs composite was attributed to the synergistic
effect between g-C3N4/Bi2O3 and N-HMCs. Moreover, the nitrogen-doped hollow
mesoporous carbon spheres (N-HMCs) were found to play a vital role in improving
the visible light absorption, charge transfer rates, and the active sites for the effec-
tive photocatalytic reactions [83]. As a concluding remark, Table 4.1 summarizes
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 107

Table 4.1 Summary of various photocatalyst systems and their degradation efficiency of


pharmaceutical pollutants
Degradation Time Light
Systems Materials Pollutants (%) (min) source Ref.
Schottky Ag/Pt@TiO2 Ciprofloxacin ~100 150 Visible [22]
PtDPA@SiO2@ Tetracycline 66 100 Visible [23]
CdS/Pt
Plasmonic Ag/Ag2MoO4 Ciprofloxacin 99.5 60 Visible [27]
Ag/Bi3O4Cl Tetracycline 94.2 120 Visible [28]
Au@ZnO Tetracycline 85.5 120 Visible [29]
BiVO4/Ag/Cu2O Tetracycline 91.22 60 Visible [31]
p-n junction CuS/BiVO4 Ciprofloxacin 86.7 90 Visible [39]
Ag2O/CeO2 Enrofloxacin 87.11 120 Visible [40]
CaFe2O4@ Norfloxacin 83 90 Solar [41]
ZnFe2O4
Ag/Ag2O/ Tetracycline 84.4 90 Visible [42]
PbBiO2Br 50.9 240 NIR
All-solid-state Ag@AgI/ Tetracycline 86.4 60 Visible [47]
Z-scheme VI-BOI
Bi5FeTi3O15/ Tetracycline ~70 60 Visible [48]
Ag/g-C3N4 ~85 20 Solar
WO3−X/Ag/ZnCr Tetracycline 92 90 Visible [49]
LDH
Ag/Ag3PO4/ Tetracycline 94.96 60 Visible [50]
BiVO4/RGO
Direct Z-scheme Sn3O4/g-C3N4 Tetracycline 72.2 120 Visible [55]
CuInS2/Bi2WO6 Tetracycline 90.5 120 Visible [57]
Fe2O3/CuBi2O4 Tetracycline ~80 120 Visible [59]
Bi3TaO7 Ciprofloxacin 91 120 Visible [60]
QDs/g-C3N4
g-C3N4@ZnO Cephalexin 98.90 60 Solar [62]
Heterojunction BiOBr/Fe3O4 Norfloxacin 90 30 Visible [64]
g-C3N4/ Tetracycline 81 90 Visible [67]
K+Ca2Nb3O10
Co2SnO4–SnO2/ Chlortetracycline ~83 80 Visible [69]
GC Tetracycline ~80
Carbon-based Ce(MoO4)2/GO Chloramphenicol 99 50 Visible [74]
hybrids TiO2–rGO Tetracycline 94 40 Visible [76]
Ag/N–GQDs/g-­ Tetracycline 92.8 60 Solar [77]
C3N4 90.11 Visible
31.3 NIR
TiO2/C-dots Levofloxacin ~100 90 Solar [78]
Carbon Tylosin 98 45 Solar [83]
spheres/g-C3N4
108 R. Karuppannan et al.

the above-discussed various photocatalyst systems and their efficiency towards the
degradation pharmaceutical pollutants. These reports from the recent works were
published during 2017–2019.

4.6  Summary and Outlook

The antibiotics are considered as an emerging pharmaceutical pollutant due to their


serious health issues to human, animals, and the ecosystem. Therefore, the complete
removal of antibiotics from the respective sources is highly important, which is also
challenging. In this direction, out of various processes, the nanocomposite-based
photocatalytic process is found to be potentially powerful and environmentally
benign towards the removal and degradation of the antibiotic molecules in water,
using the natural sunlight as a source to activate the catalysts. As the photocatalytic
properties of the materials are mostly dependent upon the photoabsorbance, charge
separation, charge transfer, surface properties, and redox potential, it is essential to
construct suitable material system to meet these requirements. In this direction, the
nanocomposite-based hybrid photocatalysts are found to be well equipped and
potential for the effective degradation of antibiotics in water medium. The further
developments in this direction involve the construction of nanocomposite photo-
catalysts for the rapid, simultaneous, and large-scale degradation of multidrug phar-
maceutical pollutants.

Acknowledgments  This work was supported by the Natural Sciences and Engineering Research
Council of Canada (NSERC) through the Collaborative Research and Development (CRD),
Strategic Project (SP), and Discovery Grants (DG). MS gratefully acknowledges the Department
of Science and Technology, Govt. of India, for the funding support through the DST-INSPIRE
Faculty Award [DST/INSPIRE/04/2016/002227, 14-02-2017].

References

1. Martínez, J.  L. (2008). Antibiotics and antibiotic resistance genes in natural environments.
Science, 321(5887), 365–367.
2. de Kraker, M. E., Stewardson, A. J., & Harbarth, S. (2016). Will 10 million people die a year
due to antimicrobial resistance by 2050? PLoS Medicine, 13(11).
3. Danner, M.-C., et  al. (2019). Antibiotic pollution in surface fresh waters: Occurrence and
effects. Science of the Total Environment.
4. Patel, M., et al. (2019). Pharmaceuticals of emerging concern in aquatic systems: Chemistry,
occurrence, effects, and removal methods. Chemical Reviews, 119(6), 3510–3673.
5. Bagheri, S., et al. (2017). Photocatalytic pathway toward degradation of environmental phar-
maceutical pollutants: Structure, kinetics and mechanism approach. Catalysis Science &
Technology, 7, 4548–4569.
6. Rosman, N., et  al. (2018). Hybrid membrane filtration-advanced oxidation processes for
removal of pharmaceutical residue. Journal of Colloid and Interface Science, 532, 236–260.
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 109

7. Marcelino, R. B. P., et al. (2017). Multistage ozone and biological treatment system for real
wastewater containing antibiotics. Journal of Environmental Management, 195, 110–116.
8. Ahmed, M. J., et al. (2018). Removal of emerging pharmaceutical contaminants by adsorption
in a fixed-bed column: A review. Ecotoxicology and Environmental Safety, 149, 257–266.
9. Mitra, N., et al. (2018). Removal of pharmaceutical compounds in water and wastewater using
fungal oxidoreductase enzymes. Environmental Pollution, 234, 190–213.
10. Calvete, M. J., et al. (2019). Hybrid materials for heterogeneous photocatalytic degradation of
antibiotics. Coordination Chemistry Reviews, 395, 63–85.
11. Li, D., & Shi, W. (2016). Recent developments in visible-light photocatalytic degradation of
antibiotics. Chinese Journal of Catalysis, 37(6), 792–799.
12. Sarkar, S., et  al. (2014). Involvement of process parameters and various modes of applica-
tion of TiO2 nanoparticles in heterogeneous photocatalysis of pharmaceutical wastes—a short
review. RSC Advances, 4(100), 57250–57266.
13. Nguyen, C. C., Vu, N. N., & Do, T.-O. (2015). Recent advances in the development of sunlight-­
driven hollow structure photocatalysts and their applications. Journal of Materials Chemistry
A, 3(36), 18345–18,359.
14. Gaya, U. I., & Abdullah, A. H. (2008). Heterogeneous photocatalytic degradation of organic
contaminants over titanium dioxide: A review of fundamentals, progress and problems.
Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 9(1), 1–12.
15. Xu, C., et al. (2019). Nanostructured materials for photocatalysis. Chemical Society Reviews,
48(14), 3868–3902.
16. Gan, Y., et al. (2018). Impact of post-processing modes of precursor on adsorption and pho-
tocatalytic capability of mesoporous TiO2 nanocrystallite aggregates towards ciprofloxacin
removal. Chemical Engineering Journal, 349, 1–16.
17. Hou, J., et  al. (2019). Narrowing the band gap of BiOCl for the hydroxyl radical genera-
tion of photocatalysis under visible light. ACS Sustainable Chemistry & Engineering, 7(19),
16569–16576.
18. Bojer, C., et al. (2017). Clinical wastewater treatment: Photochemical removal of an anionic
antibiotic (ciprofloxacin) by mesostructured high aspect ratio ZnO nanotubes. Applied
Catalysis B: Environmental, 204, 561–565.
19. Devi, L. G., & Kavitha, R. (2016). A review on plasmonic metal TiO2 composite for genera-
tion, trapping, storing and dynamic vectorial transfer of photogenerated electrons across the
Schottky junction in a photocatalytic system. Applied Surface Science, 360, 601–622.
20. Wang, P., et al. (2012). Plasmonic photocatalysts: Harvesting visible light with noble metal
nanoparticles. Physical Chemistry Chemical Physics, 14(28), 9813–9825.
21. Jiang, Z., et al. (2014). In situ synthesis of bimetallic Ag/Pt loaded single-crystalline anatase
TiO2 hollow nano-hemispheres and their improved photocatalytic properties. CrystEngComm,
16(12), 2384–2394.
22. Fang, J., et al. (2017). CdS/Pt photocatalytic activity boosted by high-energetic photons based
on efficient triplet–triplet annihilation upconversion. Applied Catalysis B: Environmental, 217,
100–107.
23. Liu, Y., Wu, Q., & Zhao, Y. (2017). Biomimetic synthesis of Ag3PO4-NPs/Cu-NWs with
visible-light-enhanced photocatalytic activity for degradation of the antibiotic ciprofloxacin.
Dalton Transactions, 46(19), 6425–6432.
24. Cai, T., et  al. (2018). Ag3PO4/Ti3C2 MXene interface materials as a Schottky catalyst with
enhanced photocatalytic activities and anti-photocorrosion performance. Applied Catalysis B:
Environmental, 239, 545–554.
25. Luo, B., et  al. (2015). Fabrication of a Ag/Bi3TaO7 plasmonic photocatalyst with enhanced
photocatalytic activity for degradation of tetracycline. ACS Applied Materials & Interfaces,
7(31), 17061–17069.
26. Li, J., Liu, F., & Li, Y. (2018). Fabrication of an Ag/Ag2MoO4 plasmonic photocatalyst with
enhanced photocatalytic performance for the degradation of ciprofloxacin. New Journal of
Chemistry, 42(14), 12054–12061.
110 R. Karuppannan et al.

27. Jiang, E., et  al. (2018). Visible-light-driven Ag/Bi3O4 Cl nanocomposite photocatalyst with
enhanced photocatalytic activity for degradation of tetracycline. RSC Advances, 8(65),
37200–37207.
28. Gao, B., et  al. (2019). Zeolitic imidazolate framework 8-derived Au@ ZnO for efficient
and robust photocatalytic degradation of tetracycline. Chinese Journal of Chemistry, 37(2),
148–154.
29. Wang, F., et al. (2018). The facile synthesis of a single atom-dispersed silver-modified ultrathin
gC3N4 hybrid for the enhanced visible-light photocatalytic degradation of sulfamethazine with
peroxymonosulfate. Dalton Transactions, 47(20), 6924–6933.
30. Deng, Y., et al. (2017). Plasmonic resonance excited dual Z-scheme BiVO4/Ag/Cu2O nano-
composite: Synthesis and mechanism for enhanced photocatalytic performance in recalcitrant
antibiotic degradation. Environmental Science: Nano, 4(7), 1494–1511.
31. Chen, Q., Wu, S., & Xin, Y. (2016). Synthesis of Au–CuS–TiO2 nanobelts photocatalyst for
efficient photocatalytic degradation of antibiotic oxytetracycline. Chemical Engineering
Journal, 302, 377–387.
32. Xue, J., et al. (2015). Au-loaded porous graphitic C3N4/graphene layered composite as a ter-
nary plasmonic photocatalyst and its visible-light photocatalytic performance. RSC Advances,
5(107), 88249–88257.
33. Li, H., et al. (2015). State-of-the-art progress in diverse heterostructured photocatalysts toward
promoting photocatalytic performance. Advanced Functional Materials, 25(7), 998–1013.
34. Li, L., Salvador, P.  A., & Rohrer, G.  S. (2014). Photocatalysts with internal electric fields.
Nanoscale, 6(1), 24–42.
35. Ao, Y., et al. (2016). Fabrication of novel p–n heterojunction BiOI/La2Ti2O7 composite pho-
tocatalysts for enhanced photocatalytic performance under visible light irradiation. Dalton
Transactions, 45(19), 7986–7997.
36. Zheng, J., & Lei, Z. (2018). Incorporation of CoO nanoparticles in 3D marigold flower-like
hierarchical architecture MnCo2O4 for highly boosting solar light photo-oxidation and reduc-
tion ability. Applied Catalysis B: Environmental, 237, 1–8.
37. Suyana, P., et al. (2017). Co3 O4–C3N4 p–n nano-heterojunctions for the simultaneous degra-
dation of a mixture of pollutants under solar irradiation. Environmental Science: Nano, 4(1),
212–221.
38. Lai, C., et al. (2019). Fabrication of CuS/BiVO4 (0 4 0) binary heterojunction photocatalysts
with enhanced photocatalytic activity for Ciprofloxacin degradation and mechanism insight.
Chemical Engineering Journal, 358, 891–902.
39. Wen, X.-J., et al. (2018). A novel Ag2O/CeO2 heterojunction photocatalysts for photocatalytic
degradation of enrofloxacin: Possible degradation pathways, mineralization activity and an in
depth mechanism insight. Applied Catalysis B: Environmental, 221, 701–714.
40. Behera, A., et al. (2019). Construction of isoenergetic band alignment between CdS QDs and
CaFe2O4@ ZnFe2O4 heterojunction: A promising ternary hybrid toward norfloxacin degrada-
tion and H2 energy production. The Journal of Physical Chemistry C, 123(28), 17112–17126.
41. Guo, H., et al. (2019). Integrating the plasmonic effect and pn heterojunction into a novel Ag/
Ag2O/PbBiO2Br photocatalyst: Broadened light absorption and accelerated charge separation
co-mediated highly efficient visible/NIR light photocatalysis. Chemical Engineering Journal,
360, 349–363.
42. Yan, M., et al. (2017). Fabrication of nitrogen doped graphene quantum dots-BiOI/MnNb2O6
pn junction photocatalysts with enhanced visible light efficiency in photocatalytic degradation
of antibiotics. Applied Catalysis B: Environmental, 202, 518–527.
43. Zhou, P., Yu, J., & Jaroniec, M. (2014). All-solid-state Z-scheme photocatalytic systems.
Advanced Materials, 26(29), 4920–4935.
44. Low, J., et  al. (2017). A review of direct Z-scheme photocatalysts. Small Methods, 1(5),
1700080.
45. Li, Q., et al. (2017). Z-scheme BiOCl-Au-CdS heterostructure with enhanced sunlight-driven
photocatalytic activity in degrading water dyes and antibiotics. ACS Sustainable Chemistry &
Engineering, 5(8), 6958–6968.
4  Nanocomposite Photocatalysts for the Degradation of Contaminants of Emerging… 111

46. Yang, Y., et  al. (2018). Construction of iodine vacancy-rich BiOI/Ag@ AgI Z-scheme het-
erojunction photocatalysts for visible-light-driven tetracycline degradation: Transformation
pathways and mechanism insight. Chemical Engineering Journal, 349, 808–821.
47. Wang, K., Li, J., & Zhang, G. (2019). Ag-bridged Z-Scheme 2D/2D Bi5FeTi3O15/g-C3N4 het-
erojunction for enhanced photocatalysis: Mediator-induced interfacial charge transfer and
mechanism insights. ACS Applied Materials & Interfaces, 11(31), 27686–27,696.
48. Sahoo, D. P., Patnaik, S., & Parida, K. (2019). Construction of a Z-Scheme dictated WO3–X/
Ag/ZnCr LDH synergistically visible light-induced photocatalyst towards tetracycline degra-
dation and H2 evolution. ACS Omega, 4(12), 14721–14741.
49. Chen, F., et al. (2017). Hierarchical assembly of graphene-bridged Ag3PO4/Ag/BiVO4 (040)
Z-scheme photocatalyst: An efficient, sustainable and heterogeneous catalyst with enhanced
visible-light photoactivity towards tetracycline degradation under visible light irradiation.
Applied Catalysis B: Environmental, 200, 330–342.
50. Lu, Z., et al. (2018). Facile microwave synthesis of a Z-scheme imprinted ZnFe2O4/Ag/PEDOT
with the specific recognition ability towards improving photocatalytic activity and selectivity
for tetracycline. Chemical Engineering Journal, 337, 228–241.
51. Liu, Q., et al. (2018). 3D reduced graphene oxide aerogel-mediated Z-scheme photocatalytic
system for highly efficient solar-driven water oxidation and removal of antibiotics. Applied
Catalysis B: Environmental, 232, 562–573.
52. Wang, T., et al. (2016). Synthesis of redox-mediator-free direct Z-scheme AgI/WO3 nanocom-
posite photocatalysts for the degradation of tetracycline with enhanced photocatalytic activity.
Chemical Engineering Journal, 300, 280–290.
53. Zhu, W., et  al. (2017). Construction of WO3–gC3N4 composites as efficient photocatalysts
for pharmaceutical degradation under visible light. Catalysis Science & Technology, 7(12),
2591–2600.
54. Li, C., et al. (2018). Z-scheme mesoporous photocatalyst constructed by modification of Sn3O4
nanoclusters on g-C3N4 nanosheets with improved photocatalytic performance and mechanism
insight. Applied Catalysis B: Environmental, 238, 284–293.
55. You, Z., et  al. (2017). Preparation of g-C3N4 nanorod/InVO4 hollow sphere composite with
enhanced visible-light photocatalytic activities. Applied Catalysis B: Environmental, 213,
127–135.
56. Lu, X., et  al. (2019). The facile fabrication of novel visible-light-driven Z-scheme CuInS2/
Bi2WO6 heterojunction with intimate interface contact by in situ hydrothermal growth strategy
for extraordinary photocatalytic performance. Chemical Engineering Journal, 356, 819–829.
57. Luo, J., et  al. (2019). Rational design of Z-scheme LaFeO3/SnS2 hybrid with boosted vis-
ible light photocatalytic activity towards tetracycline degradation. Separation and Purification
Technology, 210, 417–430.
58. Li, M.-y., et al. (2018). Design of visible-light-response core–shell Fe2O3/CuBi2O4 heterojunc-
tions with enhanced photocatalytic activity towards the degradation of tetracycline: Z-scheme
photocatalytic mechanism insight. Inorganic Chemistry Frontiers, 5(12), 3148–3154.
59. Wang, K., et al. (2017). 0D/2D Z-scheme heterojunctions of bismuth tantalate quantum dots/
ultrathin g-C3N4 nanosheets for highly efficient visible light photocatalytic degradation of anti-
biotics. ACS Applied Materials & Interfaces, 9(50), 43704–43715.
60. Ma, X., et  al. (2017). 2D/2D Heterojunctions of WO3 nanosheet/K+ Ca2Nb3O10− ultrathin
nanosheet with improved charge separation efficiency for significantly boosting photocataly-
sis. Catalysis Science & Technology, 7(16), 3481–3491.
61. Li, N., et al. (2018). Z-Scheme 2D/3D g-C3N4@ ZnO with enhanced photocatalytic activity for
cephalexin oxidation under solar light. Chemical Engineering Journal, 352, 412–422.
62. Low, J., et al. (2017). Heterojunction photocatalysts. Advanced Materials, 29(20), 1601694.
63. Guo, C., et al. (2017). Assessing the photocatalytic transformation of norfloxacin by BiOBr/
iron oxides hybrid photocatalyst: Kinetics, intermediates, and influencing factors. Applied
Catalysis B: Environmental, 205, 68–77.
112 R. Karuppannan et al.

64. Yuan, X., et al. (2019). In-situ synthesis of 3D microsphere-like In2S3/InVO4 heterojunction
with efficient photocatalytic activity for tetracycline degradation under visible light irradiation.
Chemical Engineering Journal, 356, 371–381.
65. Jin, Y., et  al. (2017). Construction of ultrafine TiO2 nanoparticle and SnNb2O6 nanosheet
0D/2D heterojunctions with abundant interfaces and significantly improved photocatalytic
activity. Catalysis Science & Technology, 7(11), 2308–2317.
66. Jiang, D., et al. (2017). Perovskite oxide ultrathin nanosheets/g-C3N4 2D-2D heterojunction
photocatalysts with significantly enhanced photocatalytic activity towards the photodegrada-
tion of tetracycline. Applied Catalysis B: Environmental, 201, 617–628.
67. Jiang, Z., et al. (2016). Constructing graphite-like carbon nitride modified hierarchical yolk–
shell TiO2 spheres for water pollution treatment and hydrogen production. Journal of Materials
Chemistry A, 4(5), 1806–1818.
68. Zheng, J., & Zhang, L. (2018). Rational design and fabrication of multifunctional catalyzer
CO2SnO4-SnO2/GC for catalysis applications: Photocatalytic degradation/catalytic reduction
of organic pollutants. Applied Catalysis B: Environmental, 231, 34–42.
69. Syed, N., et al. (2019). Carbon based nanomaterials via heterojunction serving as photocata-
lyst. Frontiers in Chemistry, 7, 713.
70. Wang, R., et al. (2017). Recent progress in carbon quantum dots: Synthesis, properties and
applications in photocatalysis. Journal of Materials Chemistry A, 5(8), 3717–3734.
71. Li, X., et al. (2016). Graphene in photocatalysis: A review. Small, 12(48), 6640–6696.
72. Xu, Y., et  al. (2018). Graphene oxide-modified LaVO 4 nanocomposites with enhanced
photocatalytic degradation efficiency of antibiotics. Inorganic Chemistry Frontiers, 5(11),
2818–2828.
73. Karthik, R., et  al. (2017). A study of electrocatalytic and photocatalytic activity of cerium
molybdate nanocubes decorated graphene oxide for the sensing and degradation of antibiotic
drug chloramphenicol. ACS Applied Materials & Interfaces, 9(7), 6547–6559.
74. Tang, Y., et  al. (2015). Enhanced photocatalytic degradation of tetracycline antibiotics by
reduced graphene oxide–CdS/ZnS heterostructure photocatalysts. New Journal of Chemistry,
39(7), 5150–5160.
75. Zhang, S., et al. (2017). Interfacial growth of TiO2-rGO composite by pickering emulsion for
photocatalytic degradation. Langmuir, 33(20), 5015–5024.
76. Deng, Y., et  al. (2017). Construction of plasmonic Ag and nitrogen-doped graphene quan-
tum dots codecorated ultrathin graphitic carbon nitride nanosheet composites with enhanced
photocatalytic activity: Full-spectrum response ability and mechanism insight. ACS Applied
Materials & Interfaces, 9(49), 42816–42828.
77. Sharma, S., et al. (2018). Solar light driven photocatalytic degradation of levofloxacin using
TiO2/carbon-dot nanocomposites. New Journal of Chemistry, 42(9), 7445–7456.
78. Di, J., et al. (2015). The synergistic role of carbon quantum dots for the improved photocata-
lytic performance of Bi2MoO6. Nanoscale, 7(26), 11433–11443.
79. Di, J., et al. (2016). Carbon quantum dots induced ultrasmall BiOI nanosheets with assembled
hollow structures for broad spectrum photocatalytic activity and mechanism insight. Langmuir,
32(8), 2075–2084.
80. Li, J., et al. (2017). Fast electron transfer and enhanced visible light photocatalytic activity
using multi-dimensional components of carbon quantum dots@ 3D daisy-like In2S3/single-­
wall carbon nanotubes. Applied Catalysis B: Environmental, 204, 224–238.
81. Wu, H., et  al. (2017). Anchoring titanium dioxide on carbon spheres for high-performance
visible light photocatalysis. Applied Catalysis B: Environmental, 207, 255–266.
82. Guo, X., et  al. (2018). Highly efficient degradation toward Tylosin in the aqueous solution
by carbon spheres/g-C3N4 composites under simulated sunlight irradiation. ACS Sustainable
Chemistry & Engineering, 6(10), 12776–12786.
83. Shao, B., et al. (2018). Nitrogen-doped hollow mesoporous carbon spheres modified g-C3N4/
Bi2O3 direct dual semiconductor photocatalytic system with enhanced antibiotics degradation
under visible light. ACS Sustainable Chemistry & Engineering, 6(12), 16424–16436.
Chapter 5
Sunlight-Mediated Plasmonic
Photocatalysis: Mechanism and Material
Prospects

Durgalakshmi Dhinasekaran, M. R. Ashwin Kishore,
and Mohanraj Jagannathan

5.1  Solar Spectrum and Photocatalysis

The radiation of the sun is the only source of energy that drives the climate system
in the universe. Nearly 50% of the sun’s radiation is in the visible region of the
electromagnetic spectrum (Fig.  5.1). The other parts are in the region of near-­
infrared, with a few from the ultraviolet spectral region. Nearly 342  W of solar
radiation receives a square meter of the spherical outer surface of the earth. In that,
around 31% is instantaneously reflected back into space. This reflection is mainly
due to the clouds, the earth’s surface, and the atmosphere. The surface of the earth
receives the radiation as a form of heat from the atmosphere, partly as infrared
radiation. This energy exchange between the atmosphere and the earth’s surface
maintains the global temperature [1–3].
The photocatalytic process begins with the absorption of light by the photocata-
lytic material, which is the ultimate factor that influences the overall efficiency [4,
5]. The first and prime factor for choosing the photocatalytic material is that the
semiconductor material should have wide bandgap energy, capable to generate
charge carriers upon absorption of incident light and promote oxidation–reduction
reactions with the creation of active radicals. These processes will help to convert
light energy into unusual chemical energy in the form of energy and environmental
cleaning applications. The advancement in this field by altering the absorbing light
radiation from UV to visible and even infrared region could develop the photocata-
lyst from the viewpoint of harvesting solar energy [6–11].

D. Dhinasekaran (*) · M. Jagannathan


Department of Medical Physics, Anna University, CEG Campus, Chennai, Tamil Nadu, India
M. R. A. Kishore
Department of Chemical Engineering, University of Seoul, Seoul, Republic of Korea

© Springer Nature Switzerland AG 2021 113


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_5
114 D. Dhinasekaran et al.

Fig. 5.1  Pictorial representation of solar spectrum

5.2  Concepts of Photocatalysis

The photogenerated oxidation-reduction reactions were first encouraged by


Fujishima and Honda in [12] stating that when applying a small electrochemical
bias, water could be split into hydrogen and oxygen upon enlightening the elec-
trodes of TiO2 single crystals. This experimental process helps to develop the com-
bustible hydrogen fuel production from the water, as a way of the solar energy
conversion process [12]. Researchers are trying to make a novel redox reaction in
the combination of inorganic and organic materials, thus helping to alter the band
edge positions [13, 14], implementing different forms of semiconductors [15–19],
integration of metals [20–23], cluster [24–29], and other allotropes of carbon in a
suitable semiconductor [30–32]. This type of modification in the material synthesis
aspect provides a route to increase efficacy of redox reactions, which has become
more interesting recently.
Among all the semiconductors, titanium dioxide has commercially proven mate-
rial for photocatalytic applications by treating polluted water and air in our environ-
ment [33–35]. It has also been used as an antibacterial and antifungal [36–38], for
removal of rotten odor [39], as nitrogen fixation [40–42], as hydrogen fuel produc-
tion [43, 44], and for cleaning up oil spills in the ocean [45, 46]. The semiconductor
material with an occupied valence band and an unoccupied conduction band can act
as a light sensitized for redox reactions. The mechanism for the photocatalytic activ-
ity of a semiconductor then starts: when a photon of light with energy matches or
exceeds the bandgap energy of the particular material, an electron is moved from the
valence band to the conduction band, and the positively charged holes left the
valence band (Fig. 5.2). The photogenerated electrons in the conduction band and
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 115

6 Ox
e- 4
Ti
Ox-
2 1

HO Ti Red+ CO2, CI-, H+ , H2O


h+ 3 5
7
Red

Fig. 5.2  Pictorial representation of charge carrier pathway in the photoelectrochemical reaction:
(1) generation of charge carriers with the help of light energy; (2) electron–hole pair recombination
to release heat; (3) commencement of an oxidative way by a positively charged hole in the valence
band; (4) beginning of a reductive route by an electron in the conduction band; (5) photocatalytic
activity to degrade the products; (6) trapping of electrons; (7) trapping of holes [53]

the positive holes in the valence band can recombine and disperse the energy in the
form of heat and also get trapped in the surface of semiconductors. This kind of
electron–hole pair recombination hinders the creation of active radicals and ends up
with low photocatalytic efficiency. Modification of materials in the form of creating
surface defects, or the introduction of the host or the development of hierarchical
semiconducting photocatalyst, which leads to avoiding the recombination rate of
electron–hole pairs and thus enhances the photocatalytic activity by drastically
increases the redox reaction. Those modifications of material should engineer the
band edge potential for the better photocatalytic performances. The valence band
potential matches or equals from +1.0 to +3.5 V vs. normal hydrogen electrodes and
conduction band electrons from +0.5 to –1.5 V vs. normal hydrogen electrodes [10,
34, 35, 47–52].

5.3  Localized Surface Plasmon Resonance

Plasmonic photocatalysis makes use of noble metal nanoparticles dispersed into


semiconductor photocatalyst and possesses two prominent features—a Schottky
junction and localized surface plasmonic resonance (LSPR). The LSPR is a distinc-
tive technique, which is able to broaden the absorption range of the visible light
spectrum in the plasmonic material. Therefore, the LSPR effect highropes the utili-
zation of the sunlight spectrum, which includes a significant percentage (≈43%) of
these wavelengths. When the photons incident on the surface of metal nanoparticles,
which occurs internal oscillation of the conduction electrons. Such resonance of the
electrons enables the photocatalyst to focus the light energy neighboring it and
guides to excellent development and commencement of electron association within
the metal nanoparticles and semiconductor material. When a small spherical metal-
lic nanoparticle is irradiated with light, the oscillating electric field causes the con-
duction electrons to oscillate coherently. This is schematically pictured in Fig. 5.3.
116 D. Dhinasekaran et al.

Fig. 5.3  Schematic of


plasmon oscillation for a
sphere, showing the
displacement of the
conduction electron charge
cloud relative to the
nuclei [54]

When the electron cloud is displaced relative to the nuclei, a restoring force arises
from Coulomb attraction between electrons and nuclei that result in the oscillation
of the electron cloud relative to the nuclear framework. The oscillation frequency is
determined by four factors: the density of electrons, the effective electron mass, and
the shape and size of the charge distribution. The collective oscillation of the elec-
trons is called the dipole plasmon resonance of the particle (sometimes denoted
“dipole particle plasmon resonance” to distinguish from plasmon excitation that can
occur in bulk metal or metal surfaces) [54].
In the early twentieth century, a pioneering work involving the optical properties
of colloidal metals by Gustav Mie provides a great insight into the term surface
plasmon (SP) [55]. Nevertheless, the term surface plasmon has been used in the
very earlier period, specifically, nanoparticle-based coloring agent in the glasses.
The making and coloring of glasses with copper metal nanoparticles in the late
Bronze Age (1000–1200 BC) was found in Fratessina di Rovigo, Italy [56]. The
most common example of nanoparticle-based coloring agent is used in the famous
Lycurgus Cup (400 AC), which exhibits various colors upon illumination inside and
outside of the cup. This is one example of surface plasmon that occurs in the metal-
lic nanoparticles [57]. To obtain a red color mosaic title, Romans have used metallic
nanoparticles [58]. For making blue color paint, the Mayan civilization has used
gold and silver nanoparticles [59]. There is a lot more evidence being discovered in
the ancient building: sculpture, glasses, pots, and other utensils for the usage of
metallic nanoparticles by tuning the colors of the products.
Our forefathers have achieved all these developments without knowing the sci-
ence behind the optical properties of metallic nanoparticles by varying different
colors while interacting with light and glasses. After the discovery of surface plas-
mon into Mie’s work, the origin of the optical properties of metallic nanoparticles
was understood. More in-depth studies have been developed to achieve more poten-
tial properties of surface plasmons for real-time applications [60].
In the development of the photocatalytic field, surface plasmon properties also
help to enhance the catalytic behavior and reduce the cost of the photocatalytic
processes by a new mechanism. Integrating metallic nanoparticles with semicon-
ducting photocatalyst creates a heterojunction between them and thus helps to
transfer the charge carrier by surface plasmon resonance-induced radiative trans-
fer processes. This mechanism has been proposed by Linic’s group [4]. This
radiative transfer of charge carriers between the semiconductor–metal hetero-
junction reduces the electron–hole pair recombination and thus enhances the
photocatalytic efficiency.
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 117

Fig. 5.4  Schematic of the E0


Schottky barrier [63]

φm
φs
CB

Ef
Ef

Electron

Hole VB

Metal Semiconductor (n-type)

5.4  Schottky Junction Barrier

One more interesting feature of plasmonic photocatalysts is that they also perform
as an electron trap. The integration of metal nanoparticles with semiconductors in
the arrangement of Schottky heterojunction provides this feature. The formation of
heterojunction helps to hinder the electron–hole pair recombination by trapping the
electrons. The term space-charge separation region or also known as the Schottky
barrier is one of the effective approaches to construct the metal-semiconductor
interfaces. At this point, the flow of electrons from one material to another can avoid
the recombination rate by altering the Fermi level as well. Commonly, during the
formation of heterojunction between the n-type semiconductor and noble metal
nanoparticles, the level of work function in the metal is higher than that of the
n-type semiconductor, and electrons will flow from the semiconductor into the
metal to adjust the Fermi energy levels (Fig. 5.4). The formation of Schottky barrier
can provide as a proficient electron trap preventing electron–hole recombination in
photocatalysis processes, which end up with excellent photocatalytic performance
[7, 61–63].

5.5  T
 heoretical Insights into Band Structure Engineering
in Semiconductors for Plasmonic Photocatalysts

Over the past decade, the development in the theoretical field to predict and under-
stand the properties of materials has undergone a drastic improvement. Density
functional theory (DFT) has been a great tool to evaluate the properties with suffi-
cient accuracy, and the overall agreement between experiments and theories is sur-
prisingly good. In DFT, using appropriate exchange-correlation function for
118 D. Dhinasekaran et al.

computing band structure is crucial because the accurate prediction of electronic


bandgaps in semiconductor compounds and insulators is one of the most important
criteria in energy research fields such as photovoltaics and photocatalysis [64–67]
Notably, the factors that predominantly determine the photocatalytic properties
are the effective mass of charge carriers, mobility, and lifetime. All these depend
mainly on the positions and shape of the valence and conduction bands in the
Brillouin zone. In semiconductors, the chemical composition and structure are the
primary factors that determine the bandgap. In most of the oxidized materials,
p-orbitals are of anion from the valence band maximum (VBM), whereas s/d states
of the cation from the conduction band minimum (CBM), as shown in Fig. 5.5. The
large energy difference between the occupied anion p -orbitals and the unoccupied
cation s/d orbitals is the reason that the oxide materials possess large bandgaps. In
order to modulate the bandgap by doping, one has to choose in appropriate fractions
in such a way that the anions with higher p-orbitals and the cations with lower
s/d -orbitals [68–70].
Not all materials display great properties for photocatalysis. In most cases, one
has to engineer the existing semiconductors to improve its photocatalytic properties.
For example, anatase TiO2 is one of the widely studied materials due to its excellent
photocatalytic activity and its other intrinsic properties [71]. However, it suffers
from poor absorption of sunlight owning to its large bandgap of 3.2 eV that results
in poor photo-conversion efficiency. A lot of efforts have been made toward the
bandgap reduction to efficiently absorb and convert visible light into chemical
energy: to name a few, monodoping of impurities (N or C) [72, 73], codoping [74,
75], heterojunction [76, 77], and graphene-based composites [78–80]. Recent
reports have proposed and demonstrated the use of noble metal nanoparticle, such
as Ag or Au in the semiconductors, that can improve the photocatalytic activity due
to the localized surface plasmon resonance (LSPR) of noble metals.
The electronic structures of Ag-doped rutile and anatase TiO2 are studied by
Gang et al. using DFT calculations. They found that the doped Ag at Ti site in rutile
and anatase TiO2 formed new intermediate bands in the bandgap region. The partial
density of states analysis indicates that these new intermediate bands originated
from the orbitals of Ag 4d states mixed with the Ti 3d states. These result in a
reduced bandgap in TiO2 and contributed to the more abundant visible light
­absorption [82]. Khan et al. studied the effect of Ag doping concentration ranging
from 2.08 to 8.33 % of the electronic and optical properties of anatase TiO2. Their
results indicate that when the doping concentration increases, the visible light
absorption enhances up to 6.25 % of Ag doping concentration, and then it decreases

Fig. 5.5  Schematic energy diagrams for


­semiconductors with VBM being anion p states
and CBM being cation s/d states
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 119

with further increase in the doping concentration. This enhancement in visible light
absorption of 6.25 % Ag-doped TiO2 is due to the optimal doping concentration and
the presence of widely distributed Ag 4d states above the valence band maxi-
mum [83].
In addition to Ag and Au, Cu is also a plasmonic metal, which exhibits LSPR. In
another report, Guo et al. have studied the influence of Cu, Ag, and Au dopants on
the electronic and optical properties of anatase TiO2. The Au and Ag incorporation
in TiO2 introduced new band states in the bandgap region, whereas Cu dopant leads
to the enhancement of d states near the uppermost of the valence band as can be
seen in Fig. 5.6a, b. Figure 5.6c shows the optical transition mechanisms of Cu-,
Ag-, and Au-doped TiO2. The bands originated by Cu 3d states near the top of the
valance band enhance the visible absorption in the range of 400–1000 nm through
Cu 3d–Ti 3d optical transition. In the case of Ag and Au, the optical transition of Ag
4d states in the conduction band and from the valence band to Au doping states,
respectively, enhances the visible light absorption. Therefore, the enhancement in
the optical absorption by the incorporation of Ag, Au, and Cu in TiO2 is due to the
incident photon frequency, the collective excitations of the conduction electrons of
noble metals are resonant, and these materials can be a good plasmonic photocata-
lyst [81].
Interestingly, recent reports have shown that hydrogen doping in a noble metal-­
free semiconductor can also exhibit LSPR. Hydrogen-doped black titania with a
core/shell structure (TiO2@TiO2−xHx) was successfully synthesized by Wang et al.
[84]. The synthesized H-doped black titania possesses the largest solar absorption
of ≈ 83% which is attributed to the amorphous shell. The LSPR is the reason for this
enhancement in the light absorption similar to Ag or Pt-loaded TiO2.

Fig. 5.6 (a) Density of states of pure and Cu-, Ag-, and Au-doped TiO2. (b) Partial density of
states of doping metals. (c) Optical transition mechanisms of Cu-, Ag-, and Au-doped TiO2 [81]
120 D. Dhinasekaran et al.

Fig. 5.7  Crystal structure of pure and hydrogenated TiO2, the localized density of states and band
structure of pure and hydrogenated TiO2 [85]

Theoretically, Ma et al. demonstrated the LSPR effect in the near-infrared range


by heavy hydrogen doping in TiO2. The idea is when TiO2 is treated with hydrogen
plasma, its electron carrier concentration increases drastically. Hence, the good
solubility and moderate donor level of hydrogen in TiO2 may tentatively attribute
the strong visible and infrared light absorption to the LSPR induced by this high
carrier concentration. They systematically investigated the solubility and donor
level of hydrogen substituted interstitially and at the oxygen site to estimate the car-
rier concentration. The near-field enhancement property through hydrogen doping
is consistent with that of noble metal nanoparticles. The authors have suggested that
similar LSPR properties can also be realized in hydrogen-doped ZnO [86]. The
improved surface-enhanced Raman scattering (SERS) is observed in TiO2 by hydro-
genation by Yang et al. It has been shown that TiO2 substrate hydrogenated for 3 h
displayed the most remarkable SERS activity and enhancement factor for R6G,
which is comparable to the Ag substrate. From the band structure and DOS for pure
and hydrogenated TiO2 shown in Fig. 5.7, it is clear that the reduction in bandgap is
due to the appearance of tailed electron energy states of Ti3+ ions at the bottom of
the conduction band by hydrogenation. In addition, the intermediate electronic level
formed close to the Fermi level due to the hydrogen doping made electron transition
easier from the valence band to the conduction band and thus significantly enhances
the light absorption capability of the TiO2. Overall, the hydrogenation in TiO2 also
exhibits LSPR, which enhances the photocatalytic efficiency [85].

5.6  Plasmonic Photocatalytic Materials

The plasmonic photocatalytic systems differ with respect to the type of interaction
between the noble metal and the semiconductor. The taxonomy of the plasmonic
photocatalytic systems together with their interaction patterns is schematically
explained in Fig.  5.8. Figure  5.8c depicts the metal nanoparticles are entirely
immersed in the semiconducting system, which is in the form of a perfect core-shell
structure. This form of the metal-semiconductor core-shell system has a great
advantage of forming a Schottky junction with efficient charge carrier mobility as
compared to partially embedded structures (Fig.  5.8b). Moreover, metal
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 121

Fig. 5.8 (a) Pure metal nanoparticles (NPs) without any semiconductor. (b) Metal NPs partially
embedded into the semiconductor and partially exposed to the environment. (c) Metal NPs having
a direct electrical contact by being fully embedded within the semiconductor without being
exposed to the environment. (d) Metal NPs isolated from the semiconductor by a non-conducting
layer to prevent direct electrical contact

nanoparticles in the core are highly protected by the semiconducting shell layer
from the agglomeration, detachment, or any dissolution characteristics of the exter-
nal environment. These benefits of the particular morphology help to increase the
absorption scattering and stability of photocatalytic reaction [87].
The noble metals like gold and silver nanoparticles are extensively used as an
LSPR probing material upon interaction with various types of semiconductors for
photocatalytic systems. Some of the criteria like morphology, size, and crystalline
nature play a vital role in the photocatalytic mechanism [88, 89].

5.6.1  Ag–Metal Oxide

Christopher et al. reported the thermo-assisted reactions involving pure Ag nanopar-


ticles, which utilizes thermal energy and low-intensity photon flux for photocata-
lytic oxidation reactions at significantly lower temperatures than the conventional
thermal processes [90]. For a high rate of hot electron injection, the energies of the
hot electrons must surpass the energy barrier. Considerable efforts are thus expanded
to theoretically study this dynamic process. For Au/TiO2, a moderate Schottky bar-
rier (SB) of approximately 1 eV is formed according to the Schottky–Mott theory.
Because the energies of hot electrons typically range between 1 eV and 4 eV above
the Fermi level, the small SB makes efficient electron transfer from Au to TiO2
energetically possible, which also renders Au/TiO2 a model system for studying and
utilizing the hot electron injection process. Due to the suitable band alignments, hot
electron injection has also been observed in many other PNP-sensitized semicon-
ductors, such as Ag/TiO2, Au/ZnO, Au/WO3, and Au/CdS (Fig.  5.9). However, a
small SB does not necessarily ensure hot electron transfer. For example, hot elec-
tron transfer has never been experimentally observed in Au/Fe2O3, although the SB
of Au/Fe2O3 is considerably smaller than that of Au/TiO2. This result indicates that
some unknown factors may also determine the hot electron transfer in addition
to the SB.
122 D. Dhinasekaran et al.

Fig. 5.9  Energy band


positions of some common
semiconductor
photocatalysts and
plasmonic metals. The
energies are given on a
potential scale vs.
vacuum [91]

Fig. 5.10  TEM cross section image of the Ag/SiO2 structures inside silica substrate over which
titania film is deposited to enhance the photocatalytic activity [92]

The existence of Ag in presence of TiO2 enhancing the photocatalytic activity by


means of localized surface plasmon in the form of layer structures of SiO2 substrate
over with Ag/SiO2 matrix and TiO2 film deposition (Fig.  5.10) was reported by
Awazu et al. in the near UV illumination region [92]. Kakuta et al. observed that Ag0
species are formed on AgBr, and AgBr is not destroyed under successive UV illumi-
nation. From the SPR phenomenon, the surrounding environment of Ag NPs is
expected to influence the property of the plasmonic photocatalysts. This provided
the idea of replacing AgCl with AgBr and led to the synthesis of the plasmonic
photocatalyst Ag@AgBr, which shows high visible light absorption. However,
Ag@AgBr has weaker visible light absorption than does Ag@AgCl, and the oxida-
tion ability of Br0 is weaker than that of Cl0. Nevertheless, Ag@AgBr was found to
be stronger photocatalytic than Ag@AgCl by a factor of 1.5. For example, Ag@
AgBr requires 10 min to completely decompose 100 mL of the 20 mg L−1 methyl
orange solution under visible light, but Ag@AgCl needs 15 min. This led us to sup-
pose that AgX particles with Ag NPs deposited on their surface might be a stable
photocatalyst under visible light and eventually to the synthesis of the first visible
light plasmonic photocatalyst Ag@AgCl (i.e., Ag–AgCl). Due to the SPR of Ag
NPs deposited on AgCl particles, the plasmonic photocatalyst Ag@AgCl shows
strong absorption in the visible region, which is almost as strong as that in the UV
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 123

region. The photo-oxidation capability of Ag@AgCl, evaluated by measuring the


decomposition of methyl orange dye, is greater than that of N–TiO2 by a factor of 8.
The repeated use of Ag@AgCl of bleaching methyl orange showed that this photo-
catalyst is stable under visible light irradiation. Furthermore, the higher photocata-
lytic efficiency of AgBr compared with that of AgCl may arise also from the fact
that AgBr has a smaller bandgap than does AgCl (2.25 vs. 3.25 eV) and absorbs
visible light. Another important observation from the study of Ag@Ag(Br,I) is that
the supporting semiconductor of Ag NPs can be used to adjust the photo-oxidation
and photo-reduction abilities of a plasmonic photocatalyst [93]. Huang et  al.
designed and fabricated a series of high-performance photocatalysts by depositing
Ag NPs on the surface of various Ag salts, AgnX (Xn_ = Cl-, Br-, I-, CrO42-,
PO4 3-,PW12O403-, SiW12O404-). The photocatalytic abilities of these catalysts Ag–
AgnX can be tuned by altering the charge of the counteranions Xn. The photocata-
lytic abilities of these photocatalysts Ag–AgnX under visible light are measured by
the decomposition of methyl orange. Ag@Ag4SiW12O40 (i.e., Ag–Ag4SiW12O40) has
the strongest photocatalytic ability among the Ag–Ag–salt catalysts; the decompo-
sition of methyl orange is complete in less than 5 min [94].

5.6.2  Au–Metal Oxide

Although the first reported example of using gold in catalysis appeared in 1925, the
chemistry of gold did not receive significant attention until the 1970s. Au NPs have
been shown to enhance UV photochemistry by serving as an electron sink for con-
duction band electrons generated in TiO2, thereby enhancing electron–hole pair
separation and lifetimes. Such charge separation is enhanced by the formation of a
Schottky junction. The first paper on Au–TiO2 for photocatalysis was reported by
Gao et al. in [88]. Due to its significant electronegativity (2.31 eV) and ionization
potential (9.22 eV) and experimental values [95], gold is a poor electron donor, and
thus it interacts only weakly with most oxides. However, those interactions can be
critically important for catalysis. Two important roles of the support include (i)
serving as a scaffold for the adsorption, nucleation, and growth of gold nanoparti-
cles and (ii) providing sites for the adsorption and activation of reagents as well as
a holding place for reaction products. The interaction of gold particles with oxide
support depends on the nature of the support, preparation method, pretreatment, and
reaction environment. In addition to the aforementioned two concepts of photoca-
talysis, there is another mechanism put forward to explain those photocatalysts in
which Au NPs deposited on insulating oxides such as zeolite, ZrO2, and SiO2 that
cannot participate in the catalysis because there is no electron transfer from the
photo-excited Au NPs to the insulator support. Au NPs have also been shown to
enable photocatalytic processes in the visible range, even for large bandgap semi-
conductors. In such photocatalyst, Au NPs are considered to be quickly heated by
absorbing visible light, thereby activating the oxidation of organic pollutants. A
clear indication that excitation of gold/support nanocomposites can drive
124 D. Dhinasekaran et al.

Fig. 5.11 (a, b) TEM


images of matchstick-like
ZnO/Au heterostructure.
(c) HRTEM image of the
ZnO nanorod. (d) HRTEM
image of ZnO/Au
heterostructure. (e–h)
STEM image and
high-resolution element
mappings of the ZnO/Au
heterostructure [97]

photocatalysis was provided by numerous studies, which demonstrated that


wavelength-­dependent reaction rates peak at the frequency where the plasmon
intensity is the highest. Since gold nanoparticles have large SPR absorption cross
sections that fall in the visible region of the solar electromagnetic spectrum, the
development of gold-based plasmonic photocatalyst appears very promising for real
applications such as environmental remediation, photocatalytic hydrogen genera-
tion, etc. [96]. One of the notable works is reported by Wu et al. [97] on the ZnO/Au
heterostructures for plasmonic photocatalytic applications (Fig. 5.11). A matchstick-­
like structure has been fabricated with ZnO as a stick with Au at the top by photo-
deposition reduction process. This structure shows enhanced plasmonic
photoelectrochemical activity and utilized for solar hydrogen production.

5.6.3  Pt–Metal Oxide

Generally, Pt can lead to a photocatalytic reduction of CO2 toward methane produc-


tion and indicate that it is the most effective co-catalyst to utilize the photogenerated
electron toward CO2 reduction. The co-deposition of noble metal islands on TiO2
has been shown to be useful in improving the efficiency of photoredox transforma-
tions, particularly when gas evolution is expected. Photocatalysis of aerobic oxida-
tion in presence of TiO2 can be successfully enhanced in the presence of Pt
nanoparticle at room temperature and under visible light condition (Shiraishi et al.
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 125

2012). It is also observed that the amount of Pt loading and the size of the Pt
nanoparticle (Fig. 5.12(i)) supported on TiO2 surface greatly influence the photo-
catalytic activity. The photocatalytic activity of aerobic oxidation of these structures
is often initiated by Pt particles in the intraband or interband transition which varies
in dark and light condition (Fig. 5.12(ii)). Metallic platinum can be deposited on
TiO2 powder by photocatalytic reduction of an aqueous suspension containing chlo-
roplatinic acid, sodium chloroplatinate, hexahydroxyplatinic acid, or platinum dini-
trodiammine. Formation of large agglomerates of small Pt particles results from
platinum’s proclivity to act as an electron accumulation center, producing a high
metal loading while the semiconductor surface remains accessible to photons and
adsorbates [98]. Ishitani et al. first systematically studied the photocatalytic reduc-
tion of CO2 in a series of metal-deposited TiO2 and found that depositing of metals
(Pd, Rh, Pt, Au, Cu2O) on TiO2 can greatly boost their photocatalytic activities for
CO2 reduction to CH4 (in decreasing order). On the other hand, the loading content
of the metal has been also an important factor since the high content of metal will
induce faster electron–hole recombination and deactivation of the photocatalyst
soon. Xie et al. studied the effect of Ag, Au, Rh, Pd, and Pt as a co-catalyst on pho-
tocatalytic behaviors of TiO2 with an emphasis on the selectivity of photogenerated
electrons for the photocatalytic CO2 reduction in the presence of H2O. It was shown
that the rate of CH4 formation increased in the sequence of Ag < Rh < Au < Pd < Pt,

Fig. 5.12 (i) (a) TEM image of Pt-supported anatase with the particle size distribution histogram
of the Pt particles. (b) HRTEM images of different particle sizes of Pt nanoparticle deposited on
anatase. (ii) Proposed mechanism of the photocatalytic performance on Pt/ anatase structure for
aerobic oxidation in (a) dark and (b) light condition. In presence of visible light, the Pt particles
are activated and (a) initiated electron transfer to anatase which (b) reduces oxygen, and
thereby generated O2– (c, d) this O2– species attracts H atom of alcohol in contact and produces the
hydroperoxide and alcoholate species
126 D. Dhinasekaran et al.

corresponding well to the increase in the efficiency of electron–hole separation. It


was pointed out that Pt is the most effective co-catalyst to extract photogenerated
electrons for CO2 reduction and the selectivity of CH4 was the highest in the pres-
ence of Pt [99]. Minoo Tasbihi et  al. concluded that, in this study, the optimum
actual amount of Pt loading is 1.4 wt.%. The enhanced photocatalytic activity of
photocatalyst can be attributed to the low-recombination rate of photogenerated
electron–hole pairs caused by the presence of Pt on the titania surface. Significantly
higher (approximately 20 times) yields of CO2 photocatalytic reduction were
achieved in the presence of photocatalyst deposited on the commercial support
compared to the powder photocatalyst [100].

5.6.4  Graphene and Noble Metals

Sunlight-derived degradation of organic pollutant is an essential and ideal approach


which possesses an important role in environmental cleaning of bioremediation of
contaminated water treatment and controlling the environmental pollution. Among
the several materials, graphene and its composites exemplified a rate of degradation
in order to compete with conventional photoactive material of TiO2 (bandgap around
3.2 eV) which doesn’t sufficiently harvest the sunlight (4%). There are several noble
metal nanomaterials that have been developed so as to compensate for the harvest
or observe the higher amount of sunlight (43%). These metal nanoparticles have the
potential ability of surface enhanced Raman spectroscopy (SERS) properties which
improves the harvest or observes the visible light by tuning their size and shape.
However, the important factor is reducing the recombination of the electron/hole
pair and increasing separation needs to be addressed, thereby enhancing the organic
pollutant degradation. In addition, the combination of graphene and noble metal
nanoparticle catalyst (plasmonic) has been designed to achieve effective charge
separation and storage from the photoexcited surface. With respect to that, visible
light plasmonic photocatalyst working mechanisms as a photogenerated electron/
hole pair at the noble metal nanoparticles to graphene interface have been explained
as well. When impinging the visible light on the noble metal nanoparticles, that
radiated electron can transfer to graphene through the interaction at the interfaces
which is due to strong SPR-induced localized electric field. This enhances the rate
of electron/hole formation followed by the photogenerated electrons and holes, that
was transferred at the interface and thereby enhancement in the photocatalytic effi-
ciency. Susanta Kumar Bhunia et al. reported that rGO–Ag nanocomposite for deg-
radation of phenol, bisphenol A, and atrazine was used to demonstrate for
photocatalytic degradation study. Herein, Ag nanoparticles act as generator and
control the electron–hole pair, while rGO surface is used for attachment of organic
pollutants. As low load concentration of Ag doesn’t produce enough reactive
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 127

oxygen species, in a similar way, there is not that much of adsorption of organic
pollutant on rGO surface covered by Ag when loading at higher concentration of
Ag, and the degradation rate determined was ~80% for 500 min with five number of
runs [101]. Hui Zhang et al. investigated that graphene sheets grafted Ag@AgCl
hybrid for RhB (Rhodamine-B) degradation. The enhancement of plasmonic photo-
catalytic degradation for Ag@AgCl/rGO hybrids shows the efficient charge separa-
tion and transfer rate compared to that of bare Ag/AgCl. Because the AgCl particle
surface is usually terminated by Cl¯ ion and negatively charged, free electrons in
metallic Ag NPs deposited over the AgCl surface were polarized because of local-
ized SPR when subjected to visible light irradiation. The plasmonic generated elec-
tron–hole moves and accumulates upon the surface of the Ag nanocrystal, while
hole to the surface of the AgCl nanoparticles oxidizes the RhB molecules. During
the irradiation of visible light, the induced electron–hole were transferred at the
RGO and Ag interface. Thus, inhibiting the reverse recombination and improves the
effective photocatalytic degradation efficiency by 95% within 16  min for Ag@
AgCL/RGO hybrid composite which enhanced fourfold compared to that of bare
Ag/AgCl [102]. Mingshan Zhu et al. expressed the graphene oxide enwrapped Ag/
AgX (X=Br, Cl) nanocomposite hybrid for visible light stable plasmonic photocata-
lyst for photodegradation of methyl orange (MO). Ag/AgX and GO constituents act
as an electron donor and electron acceptor, respectively, in the hybrid nanocompos-
ites. An induced photoelectron can flow on GO nanosheets during photocatalytic
performance (Figs. 5.13 and 5.14). Thus, charge separation/transfer can be enhanced
and the reverse recombination of electron–hole pair can be suppressed, which effec-
tively degrade the MO molecules within 120 min up to 6 cycles with ~92% degrada-
tion rate [103].

Fig. 5.13  SEM image of


Ag/AgX/GO hybrid
nanocomposites (a) Ag/
AgBr, (b) Ag/AgBr/GO,
(c) Ag/AgCl, and (d) Ag/
AgCl/GO [104]
128 D. Dhinasekaran et al.

Fig. 5.14  Photocatalytic activity of graphene


and noble metals in presence of visible light for
the degradation of methyl orange dye. (A)
Photocatalytic activities of (a) Ag/AgBr and (b)
Ag/AgBr/GO nanocomposites. (B)
Photocatalytic activities of (a) Ag/AgCl and (b)
Ag/AgCl/GO nanocomposites [104]

5.6.5  G
 raphitic Carbon Nitride (g-C3N4)-Based
Plasmonic Photocatalysis

Graphitic carbon nitride (g-C3N4) was also expected to be a potential candidate for
photocatalytic applications. This is because of its efficiency, stability, inexpensive-
ness, and semiconductor properties with the bandgap around ~2.7 eV.  Moreover,
g-C3N4 offers an inhibition of reverse recombination of charge carriers, but a lower
photocatalytic activity due to poor electrical conductivity. In order to increase the
photocatalytic efficiency of g-C3N4, it was incorporated or made a composite with
noble metals (e.g., Au, Ag, etc.) evident for a new opening in efficient visible light
responsive plasmonic nanocomposite material. The noble metal nanoparticles pos-
sess characteristic properties of surface plasmon resonance (SPR). Due to this SPR
effect which can be trapped in the sunlight, thereby improving localized electron
cloud nearby at the interface of noble metal nanoparticles/graphitic carbon nitride
(g-C3N4), Mao Ye et al. reported Ag/g-C3N4 nanosheets for visible light-driven plas-
monic photocatalyst for degradation of methyl orange (MO) dye. The photo-induced
electron–hole pair was generated under the visible light impinging on g-C3N4; these
photogenerated electrons tend to transfer to Ag nanoparticles. At the interface of
Ag/g-C3N4 due to the Schottky barrier, reduction reaction was taking place so that
separation of an electron–hole pair has occurred. The photogenerated electrons fur-
ther reduce the dissolved oxygen into superoxide radicals, and these superoxide
radicals were followed by degrading the methyl orange (MO) within 60 min up to
five cycles with ~95.2% degradation for Ag/g-C3N4 nanosheets [105]. Teng Zhou
et al. studied that Ag/AgCl/g-C3N4 composite was prepared and employed to visible
light-driven plasmonic photocatalytic degradation of methyl orange (MO) dye and
involved mechanism was explained by Z-scheme method. At the Ag nanoparticles,
plasmon-induced electrons were injected into the conduction band (CB) of AgCl.
As a result, a higher amount of electron cloud was formed on the surface of the
AgCl, which further trapped toward to form the oxygen, hydrogen radicals, and
other reactive oxygen species, while the same the g-C3N4 absorbed the visible light
which can also generate the electron–hole pairs. These electrons are transferred to
Ag nanoparticles where they could be recombined with the plasmon-induced holes
by Ag nanoparticles, while the valance band (VB) holes can oxidize the organic
substance from g-C3N4. The plasmonic photocatalyst experiments were conducted
about 3h, and degradation rate was estimated at 83.5% for Ag/AgCl/g-C3N4 [106].
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 129

5.7  Conclusion and Future Directions

The concept of plasmonic photocatalysis makes the possibility to improve the effi-
ciency of photocatalytic properties under visible light irradiation. In addition to this,
it also facilitates the prospect of using rich, abundant, sustainable sunlight sources
for energy and environmental applications such as water splitting and wastewater
treatment. In this chapter, we have discussed suitable plasmonic materials and their
composites for photocatalytic applications. Although the concept of plasmonic pho-
tocatalyst is still in its preliminary stage, in the future more research may be focused
by the research community due to its significant enhancement of photocatalytic
efficiency, which is nearly ten times larger compared to conventional metal oxides.
To further enhance the efficient utilization of surface plasmon resonance effect, the
selection of plasmonic material should overlap the absorption edge of the semicon-
ductor nanoparticles. This could be achieved by changing the shape and particle size
of the metal nanoparticles and modulating the band structure of the semiconductor
nanoparticles. With the advent of nanoscience and nanotechnology advancements,
there is still plenty of room for further advanced applied studies in this mechanism
of photocatalysis.

Acknowledgment  One of the authors, D. Durgalakshmi, acknowledges Department of Science


and Technology, India, for providing DST-INPIRE Faculty Fellowship.

References

1. Liou, K. N. (2002). An introduction to atmospheric radiation. Elsevier.


2. Sen, Z. (2008). Solar energy fundamentals and modeling techniques: atmosphere, environ-
ment, climate change and renewable energy. Springer Science & Business Media.
3. Stenflo, J. O. (2013). Solar magnetic fields: polarized radiation diagnostics. Springer Science
& Business Media.
4. Linic, S., Christopher, P., & Ingram, D. B. (2011). Plasmonic-metal nanostructures for effi-
cient conversion of solar to chemical energy. Nature Materials, 10(12), 911.
5. Ravelli, D., Dondi, D., Fagnoni, M., & Albini, A. (2009). Photocatalysis. A multi-faceted
concept for green chemistry. Chemical Society Reviews, 38(7), 1999–2011.
6. Chong, M. N., Jin, B., Chow, C. W., & Saint, C. (2010). Recent developments in photocata-
lytic water treatment technology: A review. Water Research, 44(10), 2997–3027.
7. Clavero, C. (2014). Plasmon-induced hot-electron generation at nanoparticle/metal-oxide
interfaces for photovoltaic and photocatalytic devices. Nature Photonics, 8(2), 95.
8. Kudo, A., & Miseki, Y. (2009). Heterogeneous photocatalyst materials for water splitting.
Chemical Society Reviews, 38(1), 253–278.
9. Malato, S., Fernández-Ibáñez, P., Maldonado, M.  I., Blanco, J., & Gernjak, W. (2009).
Decontamination and disinfection of water by solar photocatalysis: Recent overview and
trends. Catalysis Today, 147(1), 1–59.
10. Schneider, J., Matsuoka, M., Takeuchi, M., Zhang, J., Horiuchi, Y., Anpo, M., et al. (2014).
Understanding TiO2 photocatalysis: mechanisms and materials. Chemical Reviews, 114(19),
9919–9986.
130 D. Dhinasekaran et al.

11. Wang, Z., Liu, Y., Huang, B., Dai, Y., Lou, Z., Wang, G., et al. (2014b). Progress on extending
the light absorption spectra of photocatalysts. Physical Chemistry Chemical Physics, 16(7),
2758–2774.
12. Fujishima, A., & Honda, K. (1972). Electrochemical photolysis of water at a semiconductor
electrode. Nature, 238(5358), 37–38.
13. Kabra, K., Chaudhary, R., & Sawhney, R.  L. (2004). Treatment of hazardous organic and
inorganic compounds through aqueous-phase photocatalysis: a review. Industrial and
Engineering Chemistry Research, 43(24), 7683–7696.
14. Yan, S., Lv, S., Li, Z., & Zou, Z. (2010). Organic–inorganic composite photocatalyst of
gC3N4 and TaON with improved visible light photocatalytic activities. Journal of Dalton
Transactions, 39(6), 1488–1491.
15. Berglund, S.  P., Flaherty, D.  W., Hahn, N.  T., Bard, A.  J., & Mullins, C.  B. (2011).
Photoelectrochemical oxidation of water using nanostructured BiVO4 films. The Journal of
Physical Chemistry C, 115(9), 3794–3802.
16. Hu, X., Li, G., & Yu, J. C. J. L. (2009). Design, fabrication, and modification of nanostruc-
tured semiconductor materials for environmental and energy applications. Langmuir, 26(5),
3031–3039.
17. Huang, Z.-F., Pan, L., Zou, J.-J., Zhang, X., & Wang, L. (2014). Nanostructured bismuth
vanadate-based materials for solar-energy-driven water oxidation: a review on recent prog-
ress. Nanoscale, 6(23), 14044–14063.
18. Wang, Y., Zhang, Z., Zhu, Y., Li, Z., Vajtai, R., Ci, L., et al. (2008). Nanostructured VO2 pho-
tocatalysts for hydrogen production. ACS Nano, 2(7), 1492–1496.
19. Zhou, H., Qu, Y., Zeid, T., & Duan, X. (2012). Towards highly efficient photocatalysts using
semiconductor nanoarchitectures. Energy & Environmental Science, 5(5), 6732–6743.
20. Bhunia, K., Chandra, M., Khilari, S., Pradhan, D (2018). Bimetallic PtAu alloy nanoparticles-­
integrated g-C3N4 hybrid as an efficient photocatalyst for water-to-hydrogen conversion. ACS
Applied Materials & Interfaces, 11(1), 478–488.
21. Chen, J., Wu, X. J., Yin, L., Li, B., Hong, X., Fan, Z., et al. (2015). One-pot synthesis of
CdS nanocrystals hybridized with single-layer transition-metal dichalcogenide nanosheets
for efficient photocatalytic hydrogen evolution.Angewandte Chemie, 54(4), 1210–1214.
22. Lu, Y., Yu, H., Chen, S., Quan, X., & Zhao, H. (2012). Integrating plasmonic nanoparticles
with TiO2 photonic crystal for enhancement of visible-light-driven photocatalysis. Journal of
Environmental Science and Technology, 46(3), 1724–1730.
23. Qu, X., Brame, J., Li, Q., & Alvarez, P. J. (2012). Nanotechnology for a safe and sustain-
able water supply: Enabling integrated water treatment and reuse. Accounts of Chemical
Research, 46(3), 834–843.
24. Brimblecombe, R., Swiegers, G. F., Dismukes, G. C., & Spiccia, L. (2008). Sustained water
oxidation photocatalysis by a bioinspired manganese cluster. Angewandte Chemie, 47(38),
7335–7338.
25. Li, Q., Guo, B., Yu, J., Ran, J., Zhang, B., Yan, H., et al. (2011). Highly efficient visible-light-­
driven photocatalytic hydrogen production of CdS-cluster-decorated graphene nanosheets.
Journal of American Chemical Society, 133(28), 10878–10884.
26. Pan, L., Shen, G. Q., Zhang, J. W., Wei, X. C., Wang, L., Zou, J., et al. (2015). TiO2–ZnO
composite sphere decorated with ZnO clusters for effective charge isolation in photocataly-
sis. Industrial and Engineering Chemistry Research, 54(29), 7226–7232.
27. Yu, J., Hai, Y., & Cheng, B. (2011). Enhanced photocatalytic H2-production activity of TiO2
by Ni (OH)2 cluster modification. The Journal of Physical Chemistry C, 115(11), 4953–4958.
28. Yusuf, S., & Jiao, F. (2012). Effect of the support on the photocatalytic water oxidation activ-
ity of cobalt oxide nanoclusters. ACS Catalysis, 2(12), 2753–2760.
29. Zhao, K., Zhao, S., Qi, J., Yin, H., Gao, C., Khattak, A. M., et al. (2016). Cu2O clusters grown
on TiO2 nanoplates as efficient photocatalysts for hydrogen generation. Inorganic Chemistry
Frontiers, 3(4), 488–493.
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 131

30. Han, C., Yang, M.Q., Weng, B., & Xu, Y.J. (2014). Improving the photocatalytic activity
and anti-photocorrosion of semiconductor ZnO by coupling with versatile carbon. Physical
Chemistry Chemical Physics, 16(32), 16891–16903.
31. Tang, H., Hessel, C. M., Wang, J., Yang, N., Yu, R., Zhao, H., et al. (2014). Two-dimensional
carbon leading to new photoconversion processes. Chemical Society Reviews, 43(13),
4281–4299.
32. Zhang, Y., Tang, Z.  R., Fu, X., & Xu, Y.  J. (2010). TiO2− graphene nanocomposites for
gas-phase photocatalytic degradation of volatile aromatic pollutant: is TiO2− graphene
truly different from other TiO2− carbon composite materials? Journal of ACS Nano, 4(12),
7303–7314.
33. Khan, M. A., Akhtar, M. S., Woo, S. I., & Yang, O. B. (2008). Enhanced photoresponse under
visible light in Pt ionized TiO2 nanotube for the photocatalytic splitting of water. Journal of
Catalysis Communications, 10(1), 1–5.
34. Ni, M., Leung, M. K., Leung, D. Y., & Sumathy, K. (2007). A review and recent develop-
ments in photocatalytic water-splitting using TiO2 for hydrogen production. Renewable and
Sustainable Energy Reviews, 11(3), 401–425.
35. Tang, J., Durrant, J.  R., & Klug, D.  R. (2008). Mechanism of photocatalytic water split-
ting in TiO2. Reaction of water with photoholes, importance of charge carrier dynamics,
and evidence for four-hole chemistry. Journal of the American Chemical Society, 130(42),
13885–13891.
36. Jia, L., Qiu, J., Du, L., Li, Z., Liu, H., & Ge, S. (2017). TiO2 nanorod arrays as a photocata-
lytic coating enhanced antifungal and antibacterial efficiency of Ti substrates. Nanomedicine
(London), 12(7), 761–776.
37. Karimi, L., Yazdanshenas, M.  E., Khajavi, R., Rashidi, A., & Mirjalili, M. (2014). Using
graphene/TiO2 nanocomposite as a new route for preparation of electroconductive, self-­
cleaning, antibacterial and antifungal cotton fabric without toxicity. Journal of Cellulose,
21(5), 3813–3827.
38. Wolfrum, E. J., Huang, J., Blake, D. M., Maness, P.-C., Huang, Z., Fiest, J., et al. (2002).
Photocatalytic oxidation of bacteria, bacterial and fungal spores, and model biofilm com-
ponents to carbon dioxide on titanium dioxide-coated surfaces. Environmental Science &
Technology, 36(15), 3412–3419.
39. Antonopoulou, M., & Konstantinou, I. (2015). TiO2 photocatalysis of 2-isopropyl-3-methoxy
pyrazine taste and odor compound in aqueous phase: Kinetics, degradation pathways and
toxicity evaluation.Catalysis today, 240, 22–29.
40. Comer, B. M., Medford, A. J., (2018). Analysis of photocatalytic nitrogen fixation on rutile
TiO2 (110). ACS Sustainable Chemistry & Engineering, 6(4), 4648–4660.
41. Wang, S., Hai, X., Ding, X., Chang, K., Xiang, Y., Meng, X., et al. (2017). Light-switchable
oxygen vacancies in ultrafine Bi5O7Br nanotubes for boosting solar-driven nitrogen fixation
in pure water. Advanced Materials, 29(31), 1701774.
42. Zhao, W., Zhang, J., Zhu, X., Zhang, M., Tang, J., Tan, M., et al. (2014). Enhanced nitrogen
photofixation on Fe-doped TiO2 with highly exposed (1 0 1) facets in the presence of ethanol
as scavenger. Applied Catalysis B: Environmental, 144, 468–477.
43. Reisner, E., Powell, D. J., Cavazza, C., Fontecilla-Camps, J. C., & Armstrong, F. A. (2009).
Visible light-driven H2 production by hydrogenases attached to dye-sensitized TiO2 nanopar-
ticles. Journal of the American Chemical Society, 131(51), 18457–18466.
44. Seger, B., Pedersen, T., Laursen, A. B., Vesborg, P. C., Hansen, O., & Chorkendorff, I. (2013).
Using TiO2 as a conductive protective layer for photocathodic H2 evolution. Journal of the
American Chemical Society, 135(3), 1057–1064.
45. Li, J., Yan, L., Hu, W., Li, D., Zha, F., Lei, Z., et al. (2016). Facile fabrication of underwater
superoleophobic TiO2 coated mesh for highly efficient oil/water separation. Colloids and
Surfaces A: Physicochemical and Engineering ASPE, 489, 441–446.
46. Wei, Q., Oribayo, O., Feng, X., Rempel, G.  L., Pan, Q.  J. I., & Research, E.  C. (2018).
Synthesis of polyurethane foams loaded with TiO2 nanoparticles and their modification for
132 D. Dhinasekaran et al.

enhanced performance in oil spill cleanup. Industrial & Engineering Chemistry Research,
57(27), 8918–8926.
47. Friedmann, D., Mendive, C., & Bahnemann, D. (2010). TiO2 for water treatment: parameters
affecting the kinetics and mechanisms of photocatalysis.Applied Catalysis B: Environmental,
99(3–4), 398–406.
48. Hirakawa, T., Yawata, K., & Nosaka, Y. (2007). Photocatalytic reactivity for O2− and OH
radical formation in anatase and rutile TiO2 suspension as the effect of H2O2 addition.Applied
Catalysis A: General, 325(1), 105–111.
49. Kumar, S.  G., & Devi, L.  G. (2011). Review on modified TiO2 photocatalysis under UV/
visible light: Selected results and related mechanisms on interfacial charge carrier transfer
dynamics. The Journal of Physical Chemistry A, 115(46), 13211–13241.
50. Linsebigler, A. L., Lu, G., & Yates, J. T., (1995). Photocatalysis on TiO2 surfaces: principles,
mechanisms, and selected results. Chemical Reviews, 95(3), 735–758.
51. Nakamura, R., Tanaka, T., & Nakato, Y. J. T. (2004). Mechanism for visible light responses
in anodic photocurrents at N-doped TiO2 film electrodes. The Journal of Physical Chemistry
B, 108(30), 10617–10620.
52. Tachikawa, T., Fujitsuka, M., & Majima, T. (2007). Mechanistic insight into the TiO2 pho-
tocatalytic reactions: Design of new photocatalysts. The Journal of Physical Chemistry C,
111(14), 5259–5275.
53. Hoffmann, M. R., Martin, S. T., Choi, W., & Bahnemann, D. W. (1995). Environmental appli-
cations of semiconductor photocatalysis.Chemical reviews, 95(1), 69–96.
54. Kelly, K. L., Coronado, E., Zhao, L. L., & Schatz, G. C. (2003). The optical properties of metal
nanoparticles: The influence of size, shape, and dielectric environment. ACS Publications.
55. Mie, G. (1976). Contributions to the optics of turbid media, particularly of colloidal metal
solutions. Annalen der Physik, 25(3), 377–445.
56. Angelini, I., Artioli, G., Bellintani, P., Diella, V., Gemmi, M., Polla, A., et  al. (2004).
Chemical analyses of bronze age glasses from Frattesina di Rovigo, northern Italy. Journal of
Archaeological Science, 31(8), 1175–1184.
57. Freestone, I., Meeks, N., Sax, M., & Higgitt, C. (2007). The Lycurgus cup—A roman nano-
technology. Gold bulletin, 40(4), 270–277.
58. Mody, V. V., Siwale, R., Singh, A., & Mody, H. R. (2010). Introduction to metallic nanopar-
ticles. Journal of Pharmacy and Bioallied Sciences, 2(4), 282.
59. Casanova-González, E., García-Bucio, A., Ruvalcaba-Sil, J. L., Santos-Vasquez, V., Esquivel,
B., Roldán, M. L., et al. (2012). Silver nanoparticles for SERS identification of dyes.MRS
Online Proceedings Library, 1374, 263–274.
60. García, M. A. (2011). Surface plasmons in metallic nanoparticles: fundamentals and applica-
tions. Journal of Physics D: Applied Physics, 44(28), 283001.
61. Ding, D., Liu, K., He, S., Gao, C., & Yin, Y. (2014). Ligand-exchange assisted formation of
Au/TiO2 schottky contact for visible-light photocatalysis. Nano letters, 14(11), 6731–6736.
62. Long, R., Mao, K., Gong, M., Zhou, S., Hu, J., Zhi, M., et al. (2014). Tunable oxygen activa-
tion for catalytic organic oxidation: Schottky junction versus plasmonic effects. Angewandte
Chemie International Edition, 53(12), 3205–3209.
63. Wang, H., Zhang, L., Chen, Z., Hu, J., Li, S., Wang, Z., et al. (2014a). Semiconductor het-
erojunction photocatalysts: design, construction, and photocatalytic performances. Chemical
Society Reviews, 43(15), 5234–5244.
64. Ashwin Kishore, M., & Ravindran, P. (2017). Tailoring the electronic band gap and band
edge positions in the C2N Monolayer by P and As substitution for photocatalytic water
splitting.The Journal of Physical Chemistry C, 121(40), 22216–22224.
65. Kishore, M. A., Sjåstad, A. O., & Ravindran, P. (2019). Influence of hydrogen and halogen
adsorption on the photocatalytic water splitting activity of C2N monolayer: A first-principles
study. Carbon, 141, 50–58.
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 133

66. Le Bahers, T., Rérat, M., & Sautet, P. (2014). Semiconductors used in photovoltaic and
photocatalytic devices: Assessing fundamental properties from DFT. Journal of Physical
Chemistry C, 118(12), 5997–6008.
67. Srinivasu, K., Modak, B., & Ghosh, S. K. (2014). Porous graphitic carbon nitride: A possible
metal-free photocatalyst for water splitting. The Journal of Physical Chemistry C, 118(46),
26479–26484.
68. Ahmad, S., Mahanti, S., Hoang, K., & Kanatzidis, M.  G. (2006). Ab initio studies of the
electronic structure of defects in PbTe. Physical Review B, 74(15), 155205.
69. Chen, S., Gong, X., Walsh, A., & Wei, S., (2009). Electronic structure and stability of qua-
ternary chalcogenide semiconductors derived from cation cross-substitution of II-VI and
I-III-VI 2 compounds.Physical Review B, 79(16), 165211.
70. Liao, P., & Carter, E. A. (2013). New concepts and modeling strategies to design and evaluate
photo-electro-catalysts based on transition metal oxides. Chemical Society Reviews, 42(6),
2401–2422.
71. Zhao, Z., & Liu, Q. (2007). Mechanism of higher photocatalytic activity of anatase TiO2
doped with nitrogen under visible-light irradiation from density functional theory calcula-
tion. Journal of Physics D: Applied Physics, 41(2), 025105.
72. Shi, A. J., Li, B. X., Wan, C. R., Leng, D. C., & Lei, E. Y. (2016). Hybrid density functional
studies of C-anion-doped anatase TiO2. Chemical Physics Letters, 650, 19–28.
73. Yang, K., Dai, Y., & Huang, B. (2007a). Study of the nitrogen concentration influence on
N-doped TiO2 anatase from first-principles calculations. The Journal of Physical Chemistry
C, 111(32), 12086–12090.
74. Khan, M., Xu, J., Chen, N., & Cao, W. (2012). First principle calculations of the electronic
and optical properties of pure and (Mo, N) co-doped anatase TiO2. Journal of Alloys and
Compounds, 513, 539–545.
75. Yang, K., Dai, Y., & Huang, B. (2007b). Understanding photocatalytic activity of S-and
P-doped TiO2 under visible light from first-principles. The Journal of Physical Chemistry C,
111(51), 18985–18994.
76. Kawahara, T., Konishi, Y., Tada, H., Tohge, N., Nishii, J., & Ito, S. J. (2002). A patterned TiO2
(anatase)/TiO2 (rutile) bilayer-type photocatalyst: Effect of the anatase/rutile junction on the
photocatalytic activity. Angewandte Chemie International Edition, 41(15), 2811–2813.
77. Lin, X., Xing, J., Wang, W., Shan, Z., Xu, F., & Huang, F. J. (2007). Photocatalytic activities
of heterojunction semiconductors Bi2O3/BaTiO3: A strategy for the design of efficient com-
bined photocatalysts. Journal of Physical Chemistry C, 111(49), 18288–18293.
78. Geng, W., Liu, H., & Yao, X. (2013). Enhanced photocatalytic properties of titania–graphene
nanocomposites: a density functional theory study. Journal of Physical Chemistry Chemical
Physics, 15(16), 6025–6033.
79. Huang, Q., Tian, S., Zeng, D., Wang, X., Song, W., Li, Y., et al. (2013). Enhanced photocata-
lytic activity of chemically bonded TiO2/graphene composites based on the effective interfa-
cial charge transfer through the C–Ti bond. ACS Catalysis, 3(7), 1477–1485.
80. Yang, N., Liu, Y., Wen, H., Tang, Z., Zhao, H., Li, Y., et al. (2013). Photocatalytic properties
of graphdiyne and graphene modified TiO2: From theory to experiment. ACS Nano, 7(2),
1504–1512.
81. Guo, M., & Du, J. (2012). First-principles study of electronic structures and optical properties
of Cu, Ag, and Au-doped anatase TiO2.Physica B: Condensed Matter, 407(6), 1003–1007.
82. Xing-Gang, H., An-Dong, L., Mei-Dong, H., Bin, L., & Xiao-Ling, W. (2009). First-­
principles band calculations on electronic structures of Ag-doped rutile and anatase TiO2.
Chinese Physics Letters, 26(7), 077106.
83. Khan, M., Xu, J., Chen, N., Cao, W., Usman, Z., & Khan, D. F. (2013). Effect of Ag doping
concentration on the electronic and optical properties of anatase TiO2: a DFT-based theoreti-
cal study. Research on Chemical Intermediates, 39(4), 1633–1644.
134 D. Dhinasekaran et al.

84. Wang, Z., Yang, C., Lin, T., Yin, H., Chen, P., Wan, D., et al. (2013). H-doped black titania
with very high solar absorption and excellent photocatalysis enhanced by localized surface
plasmon resonance. Advanced Functional Materials, 23(43), 5444–5450.
85. Yang, L., Peng, Y., Yang, Y., Liu, J., Li, Z., Ma, Y., et al. (2018). Green and sensitive flexible
semiconductor SERS substrates: Hydrogenated black TiO2 nanowires. ACS Applied Nano
Materials, 1(9), 4516–4527.
86. Ma, X., Dai, Y., Yu, L., & Huang, B. J. (2014). Noble-metal-free plasmonic photocatalyst:
Hydrogen doped semiconductors. Scientific Reports, 4, 3986.
87. Zhang, X., Chen, Y. L., Liu, R.-S., & Tsai, D. P. (2013). Plasmonic photocatalysis. Reports on
Progress in Physics, 76(4), 046401.
88. Gao, Y., Lee, W., Trehan, R., Kershaw, R., Dwight, K., & Wold, A. (1991). Improvement
of photocatalytic activity of titanium (IV) oxide by dispersion of Au on TiO2. Journal of
Materials Research Bulletin, 26(12), 1247–1254.
89. Sclafani, A., Palmisano, L., & Davi, E. (1991). Photocatalytic degradaton of phenol in aque-
ous polycrystalline TiO2 dispersions: the influence of Fe3+, Fe2+ and Ag+ on the reaction rate.
Journal of Photochemistry and Photobiology A: Chemistry, 56(1), 113–123.
90. Christopher, P., Xin, H., & Linic, S. (2011). Visible-light-enhanced catalytic oxidation reac-
tions on plasmonic silver nanostructures.Nature chemistry, 3(6), 467.
91. Ma, X.-C., Dai, Y., Yu, L., & Huang, B.-B. (2016). Energy transfer in plasmonic photocata-
lytic composites. Light, Science & Applications, 5(2), e16017.
92. Awazu, K., Fujimaki, M., Rockstuhl, C., Tominaga, J., Murakami, H., Ohki, Y., et al. (2008).
A plasmonic photocatalyst consisting of silver nanoparticles embedded in titanium dioxide.
Journal of the American Chemical Society, 130(5), 1676–1680.
93. Kakuta, N., Goto, N., Ohkita, H., & Mizushima, T. (1999). Silver bromide as a photocatalyst
for hydrogen generation from CH3OH/H2O solution. The Journal of Physical Chemistry B,
103(29), 5917–5919.
94. Huang, H., Li, X., Kang, Z., Liu, Y., Li, H., He, X., et  al. (2010). Tuning metal@ metal
salt photocatalytic abilities by different charged anions. Dalton Transactions, 39(44),
10593–10597.
95. Panayotov, D. A., & Morris, J. R. (2016). Surface chemistry of Au/TiO2: Thermally and pho-
tolytically activated reactions. Surface Science Reports, 71(1), 77–271.
96. Wang, P., Huang, B., Dai, Y., & Whangbo, M.-H. J. P. C. C. P. (2012). Plasmonic photocata-
lysts: harvesting visible light with noble metal nanoparticles. Physical Chemistry Chemical
Physics, 14(28), 9813–9825.
97. Wu, M., Chen, W. J., Shen, Y. H., Huang, F. Z., Li, C. H., & Li, S. K. (2014). In situ growth
of matchlike ZnO/Au plasmonic heterostructure for enhanced photoelectrochemical water
splitting. ACS Applied Materials & Interfaces, 6(17), 15052–15060.
98. Herrmann, J.  M., Disdier, J., & Pichat, P. (1986). Photoassisted platinum deposition on
TiO2 powder using various platinum complexes.The Journal of Physical Chemistry, 90(22),
6028–6034.
99. Ishitani, O., Inoue, C., Suzuki, Y., & Ibusuki, T. J. (1993). Photocatalytic reduction of car-
bon dioxide to methane and acetic acid by an aqueous suspension of metal-deposited TiO2.
Journal of Physical Chemistry C, 72(3), 269–271.
100. Tasbihi, M., Kočí, K., Edelmannová, M., Troppova, I., Reli, M., Schomaecker, R., et  al.
(2018). Pt/TiO2 photocatalysts deposited on commercial support for photocatalytic reduction
of CO2. Journal of Photochemistry and Photobiology A: Chemistry, 366, 72–80.
101. Bhunia, S.  K., Jana, N. R, (2014). Reduced graphene oxide-silver nanoparticle composite
as visible light photocatalyst for degradation of colorless endocrine disruptors. ACS Applied
Materials & Interfaces, 6(22), 20085–20092.
102. Zhang, H., Fan, X., Quan, X., Chen, S., & Yu, H. (2011). Graphene sheets grafted Ag@ AgCl
hybrid with enhanced plasmonic photocatalytic activity under visible light. Environmental
Science and Technology, 45(13), 5731–5736.
5  Sunlight-Mediated Plasmonic Photocatalysis: Mechanism and Material Prospects 135

103. Zhu, M., Chen, P., & Liu, M. (2011). Graphene oxide enwrapped Ag/AgX (X= Br, Cl)
nanocomposite as a highly efficient visible-light plasmonic photocatalyst. ACS Nano, 5(6),
4529–4536.
104. Zhu, M., Chen, P., & Liu, M. (2012). Ag/AgBr/graphene oxide nanocomposite synthesized
via oil/water and water/oil microemulsions: a comparison of sunlight energized plasmonic
photocatalytic activity. Langmuir, 28(7), 3385–3390.
105. Ye, M., Wang, R., Shao, Y., Tian, C., Zheng, Z., Gu, X., et al. (2018). Silver nanoparticles/
graphitic carbon nitride nanosheets for improved visible-light-driven photocatalytic perfor-
mance. Journal of Photochemistry and Photobiology A: Chemistry, 351, 145–153.
106. Zhou, T., Xu, Y., Xu, H., Wang, H., Da, Z., Huang, S., et al. (2014). In situ oxidation synthesis
of visible-light-driven plasmonic photocatalyst Ag/AgCl/g-C3N4 and its activity. Ceramics
International: Part A, 40(7), 9293–9301.
Chapter 6
Photocatalytic Efficiency of Bi-Based
Aurivillius Compounds: Critical Review
and Discernment of the Factors Involved

Manjunath Shetty, Murthy Muniyappa, M. Navya Rani, Vinay Gangaraju,


Prasanna D. Shivaramu, and Dinesh Rangappa

6.1  Introduction

Severely rising environmental issues have stemmed from the increasing demand for
active environmental remediation and energy renovation techniques [1].
Photocatalysis, as a tool, is used by many researchers around the globe in address-
ing these issues. Hence, photochemical approach is progressively emerging as
potential technique in this field [2]. The photocatalytic procedures require irradia-
tion of light that generates holes and electrons from semiconductors that participate
in oxidization-reduction reactions, respectively. Overall, these redox reactions pro-
mote different photocatalytic reactions, enhancing degradation of pollutants [3],
CO2 reduction [4], H2 evolution [5], and N2 fixation [6]. It’s important to note that
in a solar energy spectrum, about 43% of energy is constituted by visible light
energy. So, it is very clear that photocatalysts should be tailor-made to be highly
responsive under visible light. The abovementioned approach has consequently
made the photocatalysis a green technique in wastewater treatment [7], fuel produc-
tion [8], and generation of electricity [9].
Many semiconductors composing of metal oxides/selenides/sulfides/phosphides/
halides/oxyhalides have been studied, for their applications as potential photocata-
lysts [10–13]. Further, metal-free materials like SiC [14], g-C3N4 [15], and Si [16]
have shown promising results. Among these, the materials with bandgap of
Eg ≥ 3 eV are referred to as wide bandgap catalysts, for example, ZnO [17] and TiO2
[18], which are very well studied and renowned for their photocatalytic ability.

M. Shetty · M. Muniyappa · V. Gangaraju · P. D. Shivaramu · D. Rangappa (*)


Department of Applied Sciences, Visvesvaraya Technological University,
Center for Postgraduate Studies, Muddenahalli, Chikkaballapur, Karnataka, India
M. Navya Rani
School of Basic and Applied Sciences, Dayanand Sagar University,
Bengaluru, Karnataka, India

© Springer Nature Switzerland AG 2021 137


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_6
138 M. Shetty et al.

Further, the materials with Eg ≤ 3 eV bandgap are called visible light bandgap semi-
conductors. For example, Si [16], BiVO4 [19], CdS [20], Bi2WO6 [21], Cu2O [22],
etc. to mention are few visible light bandgap semiconductors. Particularly, Bi-based
photocatalysts can cause an upshift of valence bands (VBs), due to hybridized O 2p
and Bi 6 s2 orbitals [23]. Moreover, Bi-based Aurivillius compounds can be excited
by visible light radiation, due to their Eg  ≤  3  eV, and this advantage has gained
major interest these days and it is very clear from the number of research publica-
tions published in recent times. Also, Bi-based visible light photocatalysts have
attracted the attention, due to stable Bi3+-containing compounds like Bi2O3 [24],
BiVO4 (Liu et al. 2020a), Bi4Ti3O12 [25], Bi12TiO20 [25], Bi2O2CO3 [26], Bi2WO6
[27], Bi2MO6 [28], BiPO4 [29], BiFeO3 [30], Bi3TiNbO9 [31], and Bi0.5K0.5TiO3
[32]. Similarly, Bi5+-containing compounds like LiBiO3 [33], NaBiO3 [34], and
KBiO3 [33] have also proved to be potential visible photocatalysts; however, their
unstable nature has made them less explored compounds till date.

6.2  Bi-Based Aurivillius Compounds

Bi-based photocatalysts are seen often in the Aurivillius series. Aurivillius phase is
represented by a series of layered structures with Bi2An  −  1BnO3n  +  1 or [Bi2O2]
[An − 1BnO3n + 1] as shown in Fig. 6.1, where A links to mono-, di-, or trivalent ele-
ments, with a coordination number of 12, and B represents a transition cation appro-
priate for an octahedron with lesser size and n represents an integer number to
reflect the number of BO6 octahedron in the perovskite layer along the direction of
packing. The [An − 1BnO3n + 1]2− block is enameled alongside the c axis in the center
of the two [Bi2O2]2+ blocks, which produce a typical two-dimensional architec-
ture [35].

Bi2O2 layers

Perovskite
layers

Bi2O2 layers

n=1 n=2 n=3

Fig. 6.1  Schematic illustration of the Aurivillius-phase structures. Adopted from ref. [36]
Copyright (2001), with permission from Elsevier
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 139

6.2.1  Bi2MO6 (M = Cr, Mo, W)

Well-known Bi-based photocatalysts with Aurivillius structural arrangement are


shown in Fig. 6.2. The orthorhombic structure of Bi2MoO6 falls into the space group
Pca2(1). The electronic configuration of BMO is theoretically simulated using den-
sity functional theory (DFT). The results simulated suggest that both the conduction
band and valence band of BMO are composed of hybridized Bi 6p, O 2p, and M nd
orbitals, where n = 3, 4, and 5 for Bi2CrO6, Bi2MoO6, and Bi2WO6, respectively. The
bandgaps expected by DFT calculations are 1.245  eV, 1.96  eV, and 2.2  eV for
Bi2CrO6 (BCO), Bi2MoO6, and BWO, respectively. The experimental results showed
the bandgaps of 2.16 eV, 2.63 eV, and 2.77 eV for Bi2CrO6, Bi2MoO6, and Bi2WO6,
respectively, which is relatively higher than DFT predicted bandgaps. However,
these results put forward indicate that Bi compounds meet the requirements to
be visible light-activated photocatalysts. Despite the narrowest bandgap, it should
be noted that Bi2CrO6 is not a proper semiconductor material for photocatalysis,
because of the smaller lifetime (0.66 ns) of photoactivated charge carriers [37]. This
has resulted in a very less study of Bi2CrO6 as an effective photocatalyst [38] .

6.2.2  (BiO)2CO3

Bismuth subcarbonate ((BiO)2CO3 or Bi2CO5) is a well-known solid carbonate with


a bandgap of 3.4  eV in the system of Bi2O3–CO2–H2O.  The (BiO)2CO3 bandgap
transition can be caused, when activated with irradiation of light wavelengths less

Fig. 6.2  Supercell of Bi2Mo6 (M = W, Mo or Cr) for the DFT calculations. Reprinted from ref.
[37] Copyright (2020), with permission from Elsevier
140 M. Shetty et al.

than 365 nm. The conduction band of (BiO)2CO3 constitutes the hybridization O 2p


and Bi 6p orbitals, while the valence band is hybridized by O 2p, Bi 6p, and C 2p
orbitals. The (BiO)2CO3 compound has displayed applications in the field photo-
catalyst, as a disinfectant, and in the oxidation of pollutants (dye, NO, etc.) in waste-
water [38].

6.2.3  Bismuth Titanate (Bi4Ti3O12)

Bismuth titanate (Bi4Ti3O12) belonging to a family of Aurivillius-phase oxides has a


structure of layered perovskite that is composed of [Bi2O2]2+ layers and [Bi2Ti3O10]2−
pseudo perovskite layers arranged interchangeably, along the c axis. Hybridization
of the valance band by Bi 6s2 orbit and O 2p orbit in Bi4Ti3O12 shows a narrow
bandgap of 2.9 eV [39]. The Bi 6S2 orbit and the O 2p orbit interaction decrease the
regularity of Bi4Ti3O12 with higher fluidity of charge improving the photoinduced
carrier’s separation. The photocatalytic reaction is carried out between layers help-
ful in separating the photoinduced carriers that are owed to the unique layered struc-
ture of Bi4Ti3O12 [38].

6.3  Mechanism of Bi-Based Materials in Photocatalysis

The organic contaminants flown into the freshwater resources are a major concern.
The degradation and mineralization of these contaminants into CO2 and H2O by a
means of photocatalytic reaction is termed as photocatalysis. This includes photo-­
generated electrons, holes, and reactive oxygen species, which sums up the photo-
catalytic reaction. The light-induced degradation of organic contaminants by using
Bi-based nanoparticles is as follows, when the photocatalyst engages in the absorp-
tion of light with an appropriate wavelength from sunlight or illuminated light
source. In the photocatalyst, the electrons are promoted from valence band to con-
duction band by absorption of photons from light energy source. As a result, posi-
tive charges (holes) and negative charges (electrons) are created on valence and
conduction bands, respectively, forming electron–hole pairs in the system.
Furthermore, the free hydroxyl radicals and superoxide radicals are formed by holes
and electrons. Here, the hydroxyl radicals are formed when holes oxidize water
molecule and hydrogen gas as a by-product, while superoxide radicals are generated
when electrons reduce oxygen molecules. These formed hydroxyls and the superox-
ide radicals attack the organic pollutants causing them to degrade into harmless
products. Prolonged recombination time for holes and electrons would further
increase the efficacy of the photocatalyst by migration to the surface of the catalyst
to degrade the adsorbed pollutants on the surface by initiating the redox reactions.
With a redox potential ranging from +1.0 to +3.5 V (vs NHE at room temperature),
the strong oxidation ability of h+ ions in the valence band is dependent on the
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 141

O2
H2O

CB
Bu
light lk r
eco
Band gap mb
ina

Su
+ tion

rfa
VB

ce
rec
om
+

bin
ati
on
light

Semiconductor
H2

H+

Fig. 6.3  Schematic diagram depicting the photocatalytic degradation mechanism. Reprinted with
permission from ref. [40], Copyright (2010) American Chemical Society

semiconductor material used and the pH of the experiment. This infers the impor-
tance of the hole in the degradation of pollutants on the surface of the photocatalyst.
Both direct and indirect oxidation can occur due to reaction with surface-bound
hydroxyl radicals (˙OH), which is through h+ ions trapped on the surface or by h+
ions on the metal surfaces. The hydroxyl radicals produced have great capability to
oxidize the contaminants and mineralize them completely. Along with this, the elec-
trons generated from photoexcitation should not accumulate to avoid excess charges
inside the catalyst particles. Achieving this would help in increasing the efficiency
of a catalyst by improving the oxidation of pollutants. If the plan is to achieve oxida-
tion of organic dyes and pollutants, then oxygen is used as a scavenger as it is easily
soluble in aqueous and other solutions at cost-effective manner. With an oxidation
potential of 2.8 V and a very short lifetime, the hydroxyl radicals are the main oxi-
dizing and reactive species. Hence, in the presence of hydroxyl radicals, the miner-
alization of substrate organic compounds and dye molecules occurs completely in a
nonselective manner. The graphical interpretation of the photocatalysis mechanism
is shown in Fig. 6.3.

6.4  Synthesis of Photocatalysts

Choosing a right synthesis procedure to achieve the desired size, morphology, and
properties associated with a photocatalyst is of paramount importance. In the case
of layered Aurivillius compounds, the properties can be widely regulated using vari-
ous dopants and by changing “n,” the layer numbers. The most widely used and
142 M. Shetty et al.

reliable approaches for the synthesis of Aurivillius compounds are, namely, hydro-
thermal approach, sol-gel method, and coprecipitation method. In this section, few
of the recent reports are reviewed, which provides the status of state-of-the-art syn-
thesis method, that are being employed in synthesis of these.

6.4.1  Hydrothermal Method/Solvothermal Method

Due to its advantages over other synthesis methods in controlling particle size [41],
morphology [42], comparatively low temperature [43], and crystallinity [44], the
hydrothermal/solvothermal synthesis is a widely used technique to synthesize
Bi-based materials. The physicochemical properties of nanomaterials that affect the
photocatalytic performance depend on the critical reaction parameters such as syn-
thesis methods used, duration of reaction, temperature, duration of heat treatment,
the surfactant used, the concentration of precursor, and pH [45]. The presence of
different solvents used in the reaction is investigated by many researchers as it is
vital in the synthesis of Bi-based products. Water, ethylene glycol, and other alco-
hols have been commonly used as solvent [46].
Jinwu bai et al. successfully investigated the degradation of ciprofloxacin using
as-prepared Bi2MoO6/reduced graphene oxide (RGO) nanocomposites as photo-
catalyst. The RGO–Bi2MoO6 nanocomposite was synthesized via hydrothermal
technique and obtained a good distribution of Bi2MoO6 nanosheets on RGO form-
ing an effective photocatalyst [47]. Kai Cai et  al. used Bi2MoO6 nanosheets for
selective oxidation of toluene to benzaldehyde under visible light and in the pres-
ence of oxygen. The role of ethylene glycol used as a solvent in the formation of
ultrathin nanosheets was investigated. The results proved the role of solvent in con-
trolling the thickness of the sheet, which had a major effect in efficient electron–
hole separation. The thinner nanosheets worked efficiently in boosting the
photocatalytic oxidation of toluene to benzaldehyde with a formation rate of
778μmol/g/h and selectivity of 96% [48]. Camila Silva Ribeiro et al. studied the
reduction of CO2 into ethanol by using Bi2MoO6 samples synthesized by varying the
pH of the solution used in the hydrothermal system. The pH of the precursor solu-
tion proved to be an important factor, which is established by the results displayed
during experiments. The best results were achieved by the samples synthesized in
the water at pH 2 and 140 °C [49].
Apart from Bi-based molybdates, Yanyan Zhao et al. focused on various mor-
phologies of Bi2WO6 nanoparticles synthesized under different pH values, tempera-
tures, and solvents. Degradation of ceftriaxone sodium at a rate of 70.18% was
achieved by ordered nanoplates forming 3D flowerlike structure, which proved to
be an outstanding photocatalyst [50]. Ai-Min Yang et al. synthesized Bi2WO6 hier-
archical microspheres with large surface area resulting in more active sites and
easier charge transportation. The higher photocatalytic activity of Bi2WO6 was
vowed to hierarchical architecture achieved by the hydrothermal approach [51].
Dongfeng Sun et  al. synthesized Bi/Bi2O2CO3 photocatalyst with heterojunction
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 143

using methanol as a carbon source as well as a solvent for one-pot hydrothermal


synthesis. The plasma-enhanced heterostructure photocatalyst proved to be efficient
in photocatalytic hydrogen production [27]. Very few reports on Bi2CrO6 nanopar-
ticles have been published due to a short lifetime of charge carriers, which is about
0.660  ns. A bandgap of 2.0  eV is the challenge faced during photocatalysis as
Bi2CrO6 nanoparticles tend to have a very high recombination rate. The hydrother-
mal approach by Zizhen Li et al. in his preliminary results showed Bi2CrO6 nanopar-
ticles can be used as photocatalyst but with very low efficiency [37].

6.4.2  Sol–Gel Synthesis

Another major approach widely studied by researchers is the sol–gel process


because of its unique advantages [52]. The sol–gel method involves formation of
discrete particles or networked polymers with an integrated network by hydrolysis
of monomer precursor in a colloidal solution. The desired material yield is obtained
once the thermal treatment is complete. The advantages of the sol–gel approach are
in its simple approach in controlling chemical composition of products at relatively
low temperatures.
Yumin Liu et al. synthesized nanoplates of Bi2WO6 using egg white protein as a
bio-template. Nanoplates obtained were of 100  nm with efficient separation of
electron–hole pairs and were owed to a high specific area and better light absorp-
tion. The combined effect of morphological and optical property exhibited better
photocatalytic activity in comparison to nanoplates of Bi2WO6 without egg albu-
min [53].
Yadan Guo et al. analyzed the synergetic effect of Bi2WO6/rectorite (BR) com-
posites synthesized by the sol-gel method. The as-prepared samples were used for
photocatalytic degradation, where it was found that the Bi2WO6/rectorite showed a
strong adsorption capability and improved degradation under visible light irradia-
tion more than 420 nm, which is clearly, in this case, due to the absorbability of
rectorite [54].
Muhammad Waqar et al. synthesized La-doped Bi2MoO6 nanomaterial success-
fully using a sol–gel approach with surfactant as sodium dodecyl sulfate. Lanthanum
doping improved the efficiency, to photodegrade the methylene blue by effectively
preventing electron-hole pair recombination [55]. Fuentes et  al. doped Bi4Ti3O12
nanoparticles with Er3+, which resulted in the decrease of lattice parameters due to
smaller ionic radius of Er3+ ions (0.88 Å) in comparison to Bi3+ ions (1.03 Å) in the
host structure. This leads to photoluminescence up-conversion with increased con-
centration of Er atoms into the lattice structure. The combination of sol–gel and
hydrothermal synthesis method was performed at 30  bar pressure and showed
enhancement in infrared emission [56]. Li et al. prepared Bi4Ti3O12/SiO2 nanocom-
posite as photocatalyst through sol–gel synthesis and achieved a total decolorization
of Reactive Brilliant Red-X3B in 90 min under the illumination of UV light [57].
144 M. Shetty et al.

6.4.3  Chemical Precipitation

Bi-based Aurivillius compounds synthesized using a homogenous aqueous solution


by chemical precipitation method have been reported recently. The coprecipitation
approach is facile and cost-effective and occurs at comparatively lower tempera-
tures than other methods. Here the synthesized particles show the lower size and
larger surface area. Lue’vano-Hipó lito et  al. developed Bi2MoO6 synthesized by
coprecipitation method assisted with ultrasound irradiation and analyzed the photo-
degradation of rhodamine B under sunlight irradiation. One hundred percent of rho-
damine B dye was degraded, after 240 min of irradiation [58].
Lavakusa Banavatu et al. studied methylene blue degradation under visible light
at room temperature with Bi2MoO6-graphite flake composite, which was synthe-
sized by coprecipitation method using pH regulation and followed by calcination.
The results showed a great improvement in the photocatalytic degradation of meth-
ylene blue, under Bi2MoO6-graphite flake nanocomposite in comparison to pure
Bi2MoO6 [59]. Ait Ahsaine et al. used the coprecipitation method for the synthesis
of lutetium doped into Bi2WO6. The highest efficiency was observed for sample
doped with 5% lutetium. The high specific surface area of Bi2WO6 nanosheets facil-
itated by lutetium doping was a major reason for efficient degradation [60]. Shuai
Li et al. modified the surface of Bi2O2CO3 using lanthanum as dopant under copre-
cipitation as well as the wet impregnation method. La3+ ions dispersed on the sur-
face of the Bi2O2CO3 ensured an effective electron scavenging to trap conduction
band electrons and decrease the electron-hole recombination in Bi2O2CO3. This
modification of Bi2O2CO3 surface, enhanced the photoinduced charge separa-
tion [61].

6.5  Morphology Control

The Bi-based Aurivillius compounds can be prepared by hydrothermal, solvother-


mal, sol–gel, coprecipitation methods, etc. Variable bandgap energies are exhibited
by Bi-based samples that were prepared under different methods with different mor-
phologies and properties. The morphology and properties of semiconductors are
related to the photocatalytic efficiency of semiconductors. Accessibility of photo-
catalysts and absorption of light can be increased by the formation of hollow and
hierarchical structures with low thickness and high specific surface area. These are
the properties that improved the migration of carriers into the surface and reduce the
dosage of photocatalyst used. Based on these facts, a photocatalyst can be designed
and built or immobilized on the substrate surface.
In hydrothermal/solvothermal synthesis method, usage of water and/or alcohol
solvents has been strongly established. In the presence of ethylene glycol, (EG), the
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 145

Solid-state
Reaction

Co-
precipitation
Bi(NO3)3

Na2MoO4 or Hydrothermal
(NH4)6Mo-O24
Method

Solvothemal
Method

Fig. 6.4  The morphologies of Bi2MoO6 nanostructures fabricated through various preparation
techniques by varying reaction conditions. Reproduced from ref. [62] with permission from the
Royal Society of Chemistry

process has been used for the synthesis of hollow nanospheres and 3D hierarchical
flowerlike Bi2MoO6 nanostructures. Gou et al. prepared template-free microspheres
of mesoporous Bi2MoO6 using the solvothermal synthesis method with EG as sol-
vent. The growth of 2D platelike structure is due to EG acting as a coordinating
agent, which induces nanoparticle assembly on the surface of the microsphere. With
the assistance of microwave heating conditions, different morphologies have been
developed [62]. The Oswald ripening via hydrothermal routes has always been an
effective formation mechanism in the development of various morphologies. Few of
the structures and morphologies of Bi2MoO6 with varying conditions reported ear-
lier have been summarized in Fig. 6.4.
Previous reports also showed that Bi2WO6 samples with different morphologies
like nanospheres, nanoplates, nanoflowers, nest shaped, and nanocages have been
sinthesized by varying surfactants, pH, templates, etc. as shown in Fig. 6.5 [63].
Morphology control in Bi2O2CO3 photocatalyst has also been studied widely
based on the hydrothermal synthesis method. As per the reports, trisodium citrate
was used to obtain different morphologies, such as sponge, plates, and hierarchical
roselike structures. Trisodium citrate has proven to be an advantage as it acts as a
carbon source along with the coordinating agents. With the varying concentrations
of 0.75 mmol and 1 mmol of trisodium citrate, the spongelike and flowerlike mor-
phologies were obtained, respectively. With the addition of 1.0 mmol ammonium
carbonate (NH4)2CO3) along with trisodium citrate, the platelike morphology was
evident as shown in Fig. 6.6 [65].
146 M. Shetty et al.

Fig. 6.5  Shape evolution of Bi2WO6 micro−/nanostructures prepared with different pH values
and/or surfactants. P123 triblockpoly(ethylene oxide)-block-poly(propylene oxide)-block-­
poly(ethylene oxide) copolymer; CTAB cetyltrimethylammonium bromide, PVP poly(vinyl pyr-
rolidone), IL ionic liquid, EG ethylene glycol. Reprinted from ref. [64]. Copyright (2011), with
permission from Elsevier

6.6  Surface Modification

In the synthesis of Bi-based photocatalysts, modification of surface has been a


widely used strategy for enhancing the photocatalytic activity. Surface modification
not only enhances activity but also augments the overall efficiency of the photocata-
lyst. Surface modification can be achieved by the use of metals, ions, polymers, and
semiconductors. If a semiconducting material is used as a modifier, the resulting
nano-sized junction would be referred to as a heterojunction. Heterojunction would
facilitate fast movements of charge carriers than in bulk junction, thanks to its large
interface ratio.
During the formation of Bi2MoO6 nanostructures, the crystal facets can be par-
ticularly stabilized to control the growth rates in different crystal planes by the use
of capping agents or surfactants. Surfactants used are usually grouped into ionic and
nonionic surfactants. Further, ionic surfactants are grouped into cationic and anionic
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 147

Fig. 6.6  The FE-SEM (a, b, d, e, g, h) and TEM (c, f, i) images of Bi2O2CO3 samples: (a–c),
spongelike sample; (d–f), roselike sample; (g–i), platelike sample. Reproduced from ref. [65] with
permission from the Royal Society of Chemistry

surfactants. For example, cetyltrimethylammonium bromide (CTAB) is a cationic


surfactant and ethylenediaminetetraacetic acid (EDTA) is an anionic surfactant.
Many reports suggest the dependence of the photocatalytic performance of
Bi-based nanostructures synthesized in the presence of surfactants. Kaiqiang Jing
et  al. synthesized CTAB-assisted Bi2MoO6 spheres formed from monolayer
nanosheets. By adjusting the time and temperature, oxygen vacancy-rich nanosheets
could be synthesized for photocatalytic oxidation of benzyl alcohol. Furthermore,
the trapping of photoexcited electrons and oxygen would further separate the pho-
togenerated electrons and holes to accelerate the activation of benzyl alcohol. The
results provided inference to prove the selective oxidation of benzyl alcohol under
visible light irradiation [66]. Jing Mu et al. synthesized Bi2MoO6 with surfactant
EDTA as a visible light-driven photocatalyst by the hydrothermal method. Due to
the smaller particle size and large surface area, Bi2MoO6 nanoparticles obtained
using EDTA were responsible for enhanced degradation of methylene blue (MB)
[67]. Chao Xu et  al. synthesized Bi2MoO6 nanoplates by hydrothermal synthesis
with PVP as a surfactant. With the effect of the PVP surfactant, the absorption band
148 M. Shetty et al.

was shifted from 500 nm to 900 nm. This would extend the activity from the UV
region to the visible region, which will improve the efficiency [68].
Peng Zhang et al. developed Bi2WO6 sphere-like nanostructures applied to the
photocatalytic degradation of RhB. CTAB was used as a surfactant, which played a
major role in the formation of lamellar-like structure. This exhibited an efficient
photocatalytic degradation of RhB [69]. Si Yue Wei et al. developed a uranyl-doped
3D hierarchical flowerlike Bi2WO6 composite with the assistance of SDS surfactant.
The higher surface area played a major factor in the efficient working of photocata-
lyst by improving durability and stability [70]. Yuanyuan Li et al. analyzed the 3D
structural changes in Bi2WO6 microspheres by varying the PVP content in the pre-
cursor. That was the significant reason for the conversion of 2D plates to 3D hierar-
chical structures, which played an important role in photocatalytic efficiency [71].
Guo-Ying Zhang et  al. have synthesized Br-doped Bi2O2CO3 nanosheets using
CTAB by hydrothermal synthesis at low temperatures. The results showed improved
degradation of RhB, because of accelerated charge separation created by {001}
facets. The bromide content in the Bi2O2CO3 nanosheets helped in charge migration
due to wideband light harvesting under simulated solar irradiation [72].

6.7  Applications of Bi-Based Aurivillius Compounds

Bi-based Aurivillius compounds have been studied for organic dye degradation,
pharmaceutical waste degradation, bacterial disinfection, H2 generation, and CO2
conversion, because of their excellent physical and chemical properties. Bi-based
Aurivillius compounds are known for their narrow bandgap with nontoxic, noncor-
rosive, and relatively low cost [73]. Some of the previous studies reported some
disadvantages such as less surface area and more recombination rate of charge car-
riers, which effectively reduce the photocatalytic activity [74]. To overcome these
problems, Bi-based Aurivillius compounds have been modified or doped with non-
metals, metals, and coating with carbon-based materials to achieve Z-scheme
heterojunction-­based photocatalysis in order to suppress the charge recombination
process [75]. Various Bi-based Aurivillius compounds have been applied for photo-
catalytic applications which are discussed in the following sections.

6.7.1  Degradation of Dye

The organic pollutants from textiles, paint, tannery, pulp, and paper industries are
the main sources of water pollution. Water pollution causes several health issues
like mutagenicity, carcinogenicity, cytotoxicity, and genotoxicity followed by fail-
ure of major organs like the brain, liver, and kidneys [76]. To treat these organic
contaminants, many conventional techniques like adsorption, electrocoagulation,
biodegradation, flocculation-coagulation, and ion exchange process are being used.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 149

However, these are not reliable processes because of additional post-process require-
ment and non-efficiency.
An advanced oxidation process like photocatalytic degradation is one of the easy,
economical, and efficient processes, in which the organic contaminants break down
into H2O and CO2 by the photocatalytic reaction. It involves the generation of elec-
trons in the conduction band and holes in the valance band simultaneously by
absorbing the appropriate wavelength of light. The holes at the valance band will
oxidize hydroxyl groups adsorbed on the surface and create super hydroxyl radicals.
The free electrons at the conduction band will reduce oxygen to superoxide radicals.
These hydroxyl and superoxide radicals are very active which break down the bond-
ing of organic compounds.
Recently many researchers are focusing on the Z-scheme photocatalytic system
to suppress the recombination. Charge recombination is the major drawback in wide
bandgap materials. In Z-scheme system, there will be two catalysts combined to
form one oxidation photocatalyst and another reduction photocatalyst. In this kind
of system, there will be a greater number of electrons in the reduction catalyst,
which helps to form more superoxide radicals, whereas holes in the oxidation pho-
tocatalyst which help to form super hydroxide radicals. A higher number of super-
oxide and hydroxide radicals help to degrade more quantity of organic dyes and,
also, possess a higher absorption range. Thus, Z-scheme-type photocatalysts are
emerging as more efficient photocatalyst in these days [77]. The charge recombina-
tion process in Z-scheme process is depicted in Fig. 6.7.
Xia studied visible light-active g-C3N4/Bi2MoO6, two-dimensional Z-scheme-­
based catalysts for degradation of rhodamine B dye. They found that the photodeg-
radation efficiency of g-C3N4/BiMoO6 was more compared to pure BiMoO6. The

Fig. 6.7  Charge separation


process in Z-scheme
process
150 M. Shetty et al.

possible reason was Z-scheme-based photocatalytic activity. Here recombination of


charge carriers is sufficiently suppressed [77]. J. Jia et.al studied novel MgFe2O4/
Bi2MoO6 heterostructures for degradation of malachite green. In this study, they
found that MgFe2O4 nanoparticles were closely anchored on the surface of the
Bi2MoO6 and formed a Z-type system, which enables more light absorption capac-
ity as well as suppression of recombination rate. The sample with 20 wt% MgFe2O4/
Bi2MoO6 heterostructures shows 8.07-fold higher photocatalytic activity than pure
Bi2MoO6 [78]. H. Liu studied the SnS2/Bi2MoO6 hybrid material for the removal of
crystal violet and elucidated the charge trapping process in the Z-scheme struc-
ture [79].
Zizhen Li studied MoS2/ Bi2MoO6 nanocomposite for the removal of RhB dye.
A very thin layer of MoS2 nanosheets was wrapped onto the microsphere of Bi2MoO6
and shows very good catalytic activity because of the formation of Z-scheme het-
erostructure photocatalytic mechanism [80]. Most of the recent studies were focused
on the synthesis of Bi-based Aurivillius compounds based on Z-scheme-type photo-
catalysts for enhanced degradation of organic dye. Some of the recent studies on
Bi-based materials for dye degradation are shown in Table 6.1.

6.7.2  Degradation of Organic (Non-Dye-Based) Pollutants

Non-dye-based organic compounds are used in agricultural-based industrial appli-


cations and preservative agents [91]. Uptake of these chemicals causes cancer and
other health issues [92]. The conventional techniques have failed to treat these car-
cinogenic organic compounds. Therefore, advanced techniques have been explored
for the treatment of carcinogenic pollutants. Among various methods, photocata-
lytic degradation is a promising method to degrade organic (non-dye-based) pollut-
ants. Bi-based Aurivillius materials have gained more attention because of matching
bandgap for visible light, high stability, and nontoxicity but still have limitations
like charge recombination and less visible light capture.
In an effort to increase the efficiency of Bi-based materials, strategies like hetero-
atomic doping, solid solutions, stoichiometric changes, introduction of vacancies/
defects, ultrathin structures, hierarchical structures, hollow and porous structures,
crystal plane design, heterojunction, and electron transport materials have been
developed. Some of the recent studies on organic pollutant degradation are dis-
cussed in the next section.
Tianjin Ma et al. synthesized novel g-C3N4/C@Bi2MoO6, Z-scheme heterojunc-
tion by hydrothermal deposition process for degradation of β-naphthol. The carbon
film acts as an electron mediator for graphitic carbon nitride and Bi2MoO6. The
g-C3N4 provides active sites as well as more surface area for adsorption of the maxi-
mum amount of β-naphthol. A 10% g-C3N4 was the optimized amount for better
photocatalytic activity [93]. Yuqi Cui et.al reported degradation of 4-chlorophenol
by Bi2O3 nanoparticle-decorated anodized TiO2 nanotube arrays. Enhanced light
absorption and low charge recombination resulted in 92.5% degradation of
4-­chlorophenols under visible light [94].
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 151

Table 6.1  Studies on degradation of dye molecules by Bi-based Aurivillius materials


Concentration of dye
Sl. Name of the photocatalyst and degradation
no Name of the dye with dosage and source percentage Reference
1 Auramine-O (AO) Bi2WO6/AgS/ZnS MG 6.08 mg/L [76]
and methyl green 8–16 mg/L and visible light AO 4.04 mg/L
(MG) MG 82.75% in 60 min
AO 77.41% in 60 min
2 Rhodamine B (RhB) Bi4O5I2 1D hollow tube RhB 10 mg/L [81]
20 mg in 100 mL dye 98% in 120 min
Visible light
MoS2/Bi2MoO6 RhB 10 mg/L [80]
UV 98% in 120 min
0.1 g in 100 mL RhB
Bi2MoO6/BiFeO3 RhB 10 mg/L [82]
Visible 68% in 5 h
0.05 g in 50 mL RhB RhB 10 mg/L [83]
90.5% in 60 min
Ag/AgCl/Bi2MoO6 RhB 10 mg/L [84]
Visible 75% in 60 min
0.05 g in 50 mL RhB
CdS/Ag/Bi2MoO6
Visible
200 mg in 200 mL RhB
3 Rhodamine B (RhB) Bi2MoO6/N-rGO RhB and MB10 ppm [85]
and methylene blue
(MB)
0.1 g in 100 mL dye RhB 84% in 90 min
Visible 0.1 g in 100 mL dye, MB 82% in 90 min
Visible are related to the
Bi2MoO6/N-rGO
gC3N4/RGO/Bi2MoO6 RhB 10 mg/L [77]
Visible RhB 75% in 120 min
Bi2MoO6/ZnSnO3 MB10 mg/mL [86]
Visible MB 95% in 60 min
BiPO4/Bi2MoO6 MB10 mg/mL
Visible MB 72% in 60 min [87]
Bi2MoO6/Ag3PO4 MB10 mg/mL
Visible MB 96.8% in 1.5 min [88]
30 mg in 30 mL MB
4 Reactive red (RR) Bi2MoO6 (low temperature RR 45% in 240 min [89]
Congo red (CR) solvothermal) CR 77% in 240 min
Visible
100 mg in 100 mL RR
5 Reactive blue 19 Bi2MoO6/Bi2WO6/MWCNTs RR 96.95% in 240 min [90]
(RB) Visible
30 mg in 50 mL RB
6 Malachite green MgFe2O4/Bi2MoO6 MG 95% in 90 min [78]
(MG) Visible
30 mg in 50 mL MG
152 M. Shetty et al.

Xilu Wu et.al reported degradation of p-nitrophenol by Bi4NbO8Cl synthesized


by the solvothermal method. During the solvothermal process, more oxygen-­
deficient structures were formed as an additional source to Bi atoms, in the crystal
structure of the Bi4NbO8Cl. The Schottky barrier has formed between Bi and
Bi4NbO8Cl, which suppresses the electron-hole recombination. This enables the
efficient degradation of p-nitrophenol [95]. Qian Yang reported the degradation of
4-nitrophenol by Bi2MoO6/TiO2 heterostructures, where TiO2 nanosphere and
Bi2MoO6 nanosheets formed porous structure with abundant active sites for a redox
reaction. Due to the formation of Z-scheme heterojunction, the recombination was
reduced, and obviously, the photocatalytic efficiency and cyclability increased [96].
Feng Fu et al. reported Bi2O3/Bi2MoO6 heterostructure-based photocatalyst for the
removal of phenol, in which 96.4% of photocatalytic degradation was observed.
The improved degradation was attributed to the Z-scheme mechanism. Bi2O3
nanosheets were grown on Bi2MoO6 three-dimensional microspheres. Hence,
formed heterojunction causes reduced e− and h+ recombination with enhanced light
absorption capacity [97]. Qianyun Ron et al. reported 99.9% degradation glycolic
acid under visible light irradiation at pH  4 for 150  min by the novel Bi2MoO6/
Bi2Ti2O7 Z-scheme heterojunction material. They studied redox reaction trapping
experiments to confirm the formation of reactive oxygen and hydroxide species and
deduced the photocatalytic mechanism [98].
Based on the previous studies on Z-scheme heterojunction, Bi-based Aurivillius
material shows remarkable photocatalytic efficiency compared to a p-n junction or
type II heterojunction semiconducting materials.

6.7.3  Degradation of Pharmaceutical Pollutants

Pharmaceutical contamination occurs mainly due to antibiotics and painkiller com-


pounds. The lack of proper treatment and disposal of these antibiotics leads to con-
taminated sewage water, which later pollutes the fertilized soil causing severe health
risks for mankind. The lists of antibiotics which are studied mostly include tetracy-
cline, ibuprofen, diclofenac, carbamazepine, ciprofloxacin, metoprolol, sulfa-
methoxazole, α-ethinyl estradiol, estriol, and norfloxacin. These complex structured
antibiotics and painkillers cannot be degraded through conventional treatment
methods like adsorption, filtration, and coagulation process. For the efficient treat-
ment of pharmaceuticals, it is necessary to adopt the most efficient techniques for
the complete removal of pharmaceutical wastes from contaminated sources. Among
such techniques, photocatalytic degradation was found to be a promising approach.
Recently different types of semiconducting-based photocatalysts were studied,
which are less efficient. Among various metal oxide-based semiconductors,
Bi-based Aurivillius compounds were applied in the photocatalytic degradation of
pharmaceuticals because of excellent electrical and optical properties. The Bi-based
Aurivillius pristine compounds such as Bi4V2O11, Bi2MoO6, Bi2WO6, and Bi4Ti3O12
show less photocatalytic activity due to electron-hole recombination and less
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 153

absorption of light. To enhance the photocatalytic activity of these, material modifi-


cation is required in terms of composite or heterostructure materials. In this regard,
many studies were reported and conducted as discussed in the following section.
Bingtao Shi reported Ag/AgCl-decorated Bi4Ti3O12 nanosheets for degradation
of carbamazepine tablets under visible light irradiation. The synergistic effect of
surface plasmon resonance of Ag atoms and [001] phase of Bi4Ti3O12 was the key
for the degradation of carbamazepine, but the stability and absorption range was
limited in the visible region. To overcome these problems, the heterojunction-based
composite materials were synthesized to achieve Z-scheme mechanism [99].
Recently, W. J. Xue et al. synthesized AgI/Bi2WO6 heterojunction photocatalysts for
degradation of tetracycline, and they found that 20 wt% AgI/Bi2WO6 heterojunction
photocatalysts show 93.05% photocatalytic degradation within 60 min because of
Z-scheme mechanism, and the same has been confirmed through radical trapping
and electron spin resonance (ESR) techniques [100]. The reaction mechanism was
explained in the following steps. Upon irradiation of the appropriate amount of light
energy on AgI/Bi2WO6, a pair of electrons and holes will be generated at the valance
band and conduction band, respectively. The electron and hole pairs are generated
at both AgI and Bi2WO6. The conduction band electrons of AgI (−0.33 eV) react
with O2 to form superoxide (.O2−) radicle, and then superoxide radicle combines
with H+ ions, generated at valance band to form hydrogen peroxide (H2O2), which
further reduced into hydroxyl radicals (˙OH). The generated reactive superoxide and
hydroxyl radicals oxidize the complex tetracycline to various smaller molecules
such as CO2, H2O, etc.

(
AgI / Bi 2 WO6 + hν → AgI / Bi 2 WO6 e − + h + )
e − + O2 → O2 −

O 2 − + 2H + + e − → H 2 O 2

H 2 O2 + e − →Ù OH + OH −

(O 2

)
+ h + , + . OH + tetracycline → degraded pollutants

T.R Bastami et al. studied magnetically active Fe3O4/Bi2WO6 hybrid nanomateri-


als for degradation of ibuprofen (IBP) under visible light. The effect of pH was also
studied for the degradation of IBP with an optimum pH of 4.7.
Wei liang reported tetracycline degradation by biomass-derived carbon-modified
Bi2WO6. The Bi2WO6/C (6:1) three-dimensional oriented, flower morphology dis-
played 85.94% degradation in 90 min. The high specific surface area and long-range
visible light absorption was the main reason for enhanced photocatalytic activ-
ity [101].
Recently, X.  Liu et  al. reported novel two-dimensional heterogeneous hybrid
material Ni(OH)2/Bi4Ti3O12 and tested for degradation of the levofloxacin antibiotic
complex. This nanocomposite material exhibits higher photocatalytic activity com-
pared to pure samples, which facilitates the separation of electron holes in relation
154 M. Shetty et al.

Table 6.2  Studies on degradation of pharmaceuticals by Bi-based Aurivillius materials


Name of the
Sl. Name of the organic photocatalyst with Concentration of dye and
no chemical dosage degradation percentage Reference
1 Carbamazepine Ag/AgCl decorated 5 mg/L [99]
Bi4Ti3O12 80% in 120 min
2 Tetracycline MBi2B2O (M = Ca, Sr) 20 mg/L [103]
hydrochloride (TH) 50 mg in 50 mL TH CaBi2B2O 70% in 4 min
300 W high-pressure SrBi2B2O 58% in 4 min
mercury lamp
3 Tetracycline Bi2WO6/C 20 mg/L [101]
100 mg in 10 mL TC 85.4% in 90 min
Visible light
4 Ofloxacin (OFL) Bi2MoO6 (low OFL 85% in 240 min [89]
Norfloxacin (NOR) temperature NOR 70% in 240 min
solvothermal)
Visible
5 Levofloxacin Ni(OH) 2/Bi4Ti3O12 70% in 80 min [102]
Visible

to interfacial interactions [102]. Therefore, this study provided a new understanding


of the design of 2D/2D nanocomposites, to improve photocatalytic activity. Some of
the studies on the degradation of pharmaceuticals by Bi-based Aurivillius materials
are shown in Table 6.2.

6.7.4  Heavy Metal Reduction

Water pollution is one of the major environmental crises globally. The main source
of water pollution is industrial wastes that contain organic dyes, chemicals, pharma-
ceutical chemicals, and heavy metals. The dangerous heavy metals include Cr (VI),
Hg (II), Fe (III), and Ni (II). Uptake of these heavy metals results in mutagenesis
and carcinogenesis, which causes severe health issues. Compared to all other heavy
metals, Cr (VI) shows heavy toxicity and is very difficult to treat as it easily dis-
solves in water. Therefore, the reduction of Cr (VI) to less toxic Cr (III) is essential.
For achieving heavy metal reduction, various metal oxides and sulfides have been
studied as photocatalytic reducing agents. However, Bi-based Aurivillius com-
pounds show significant reduction abilities for removing heavy metals.
Photocatalytic reduction of heavy metal involves excitation of e− from VB to CB
by irradiating light of wavelength equivalent to the bandgap of the semiconductor.
The excited electrons reduce chromate ion into Cr3+, whereas the generated holes
were captured by hole scavengers like methanol, citric acid, etc. The mechanism of
chromium reduction is shown in Fig. 6.8.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 155

Fig. 6.8 Photocatalytic Visible


reduction mechanism of Cr Light
(VI) heavy metal. Adapted
by permission from ref.
[104]. Copyright © 2017, Cr2O72-
Springer Nature -
e-
Cr3+ CB
e- e-
CB

Eg = 2.18eV

VB
h- h-
Eg = 3.00eV

CH3OH

VB CO2+H2O
h+

There have been many studies on Bi-based Aurivillius compounds for the reduc-
tion of heavy metals that are discussed as follows. Yang et al. reported on Bi2MoO6–
Au hybrid nanostructures that are synthesized by the hydrothermal method.
Different weight ratios of Au nanoparticles were produced for the reduction of
hexavalent chromium. The Au nanoparticles coated on the surface of the Bi2MoO6
helped to decrease the electron–hole pair recombination rate by capturing excited
electrons as well as it nullifies the conduction problem of Bi2MoO6. As expected, it
was helpful to reduce a greater number of chromate ions with simultaneous oxida-
tion of benzyl alcohol [105]. Ali Rauf et al. studied hierarchical Bi2S3/Bi2WO6 het-
erojunction nanocomposite for the reduction of hexavalent chromium ions. The
effect of citric acid as a hole scavenger in the photocatalytic process was discussed.
Citric acid to Cr (VI) ratio of 2.1 resulted in a 95% reduction capacity. When the
ratio increased to 2.63, the reduction capacity was decreased to 88%. The decrease
in reduction capacity was attributed to the covering of the functional group on the
surface of Bi2S3/Bi2WO6 nanocomposite which reduces the adsorption of chromate
ions [106]. Chol-nam Ri et al. fabricated Bi2WO6/Bi4V2O11 composite for Cr (VI)
reduction under visible light. Nanocrystals of Bi4V2O11 were coated on the Bi2WO6
flakes, which enables the formation of heterojunction with good interfacial contact
to enhance quick charge transfer. Compared to individual Bi2WO6 and Bi4V2O11, the
fabricated heterojunction Bi2WO6/Bi4V2O11 showed ten times higher photocatalytic
activity with better stability. The interfacial contact of Bi2WO6 and Bi4V2O11 was the
key to achieving a good reduction capacity [107].
Yufang Liu reported a reduction of Cr (VI) by Bi2WO6/CdS heterostructure syn-
thesized by the hydrothermal method. The Bi2WO6/CdS heterostructure shows a
99% reduction of Cr (VI) within 10 min. In Bi2WO6/CdS the difference in the band
156 M. Shetty et al.

edge potential forms a heterojunction, acting as an electron-hole separation barrier


that enables successful utilization of excited electrons in the reduction of chromium
ions [103]. Qing Meng et al. reported quantum dot-decorated Bi2MoO6/boron car-
bon nitride, two-dimensional heterostructures by solvothermal method for the
reduction of Cr (VI). The synthesized quantum dot-decorated Bi2MoO6 formed
homojunction band structure and Bi2MoO6–BCN formed heterojunction. This has a
suppressed electron-hole recombination rate being the main reason for enhanced
photocatalytic reduction of Cr (VI) [108].
Juan Yang et al. reported the Bi2WO6–rGO nanocomposite for organic dye deg-
radation and Cr (VI) ion. Reduction. The strong interfacial contact and electronic
interaction were achieved by self-assembly of rGO nanosheets on Bi2WO6. Unusual
electronic interaction between rGO and Bi2WO6 shifts energy levels into higher
potential. The increased conductivity and shift in band edge eventually helped to
reduce the charge recombination, which enhanced the reduction capacity of Cr (VI)
compared to pristine Bi2WO6 [109].

6.7.5  Antibacterial Disinfection

Water contamination by microorganisms such as bacteria and viruses is a cause of


severe health issues in the entire animal family in the world. Presently, bacterial
contaminated water sources are treated by traditional methods like chlorination,
ozonation, and UV irradiation techniques, but these techniques are not efficient as
well as more costly. Therefore, the development of efficient and cost-effective pho-
tocatalytic antibacterial disinfection methods has considerable attention.
Semiconducting materials such as TiO2 and ZnO have been employed as antibacte-
rial disinfectants, but these are wide bandgap materials that are active under UV
light. In the solar spectrum, only 5% consists of UV light and more than 40% con-
sists of visible light; therefore the development of visible light-active materials is
inevitable. Among various visible light-driven photocatalysts, Bi-based Aurivillius
materials are devoted with considerable attention to bacterial disinfection.
The bacterial disinfection mechanism involves damage to the bacterial cell wall
by reactive oxygen and hydroxyl radicals. These reactive species are formed by
redox reactions of charge carriers, which are generated by absorbing light energy,
corresponding to the bandgap of material as shown in Fig. 6.9. Generally, in bacte-
rial interruption process, gram-negative bacteria E. coli and gram-positive bacteria
S. aureus were used as model organisms for study purposes. The photocatalytic
disinfection of bacteria was studied in a cultured bacterial medium. Some of the
recently reported bacterial disinfection by Bi-based Aurivillius materials is dis-
cussed below.
Mingemi Li et al. reported plasmon-induced bacterial disinfection by modified
Bi2MoO6/Ag–AgCl photocatalyst under visible light for 30 min. The enhanced pho-
tocatalytic activity was achieved mainly because of the synergistic effect of Ag
nanoparticles present on the flowerlike Bi2MoO6. Here, Bi2MoO6 acts as an
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 157

Fig. 6.9  Mechanism of


bacterial disinfection by
photocatalyst. Adopted
from ref. [110]

effective interfacial medium for charge transfer by suppressing recombination as


well as increasing light absorption capacity. The plasmonic activity of metallic Ag
particles present on the surface of Bi2MoO6 was the added advantage for excellent
bacterial disruption [111].
Pulgarin et al. synthesized flowerlike morphology of Bi2WO6 for bacterial inac-
tivation under visible light irradiation. Further, they explored the interfacial charge
transfer process between catalyst surface and bacteria. Interestingly 2.5 times higher
activity was achieved in flower-shaped Bi2WO6 compared to particles. Furthermore,
this confirms the influence of morphology on catalytic activity in bacterial disinfec-
tion [110]. Vignesh Shanmugam et al. studied the bacterial disinfection of gram-­
positive bacteria by Bi2MoO6-Ag-coated g-C3N4 under visible light. The synthesized
Bi2MoO6–Ag-coated g-C3N4 showed effective degradation of bacteria by suppress-
ing the recombination of charge carriers as well as increased visible light absorption
[112]. Recently, Huanxian Shi et al. studied photocatalytic deactivation of E. coli
cells with novel CuBi2O4/ Bi2MoO6 heterostructures. They reported the effect of
reactive oxygen species on bacterial interruption by scavenger test. The enhanced
disruption of E. coli bacteria was due to the increased production of reactive oxygen
species and higher charge separation process. The synthesized CuBi2O4/Bi2MoO6
heterostructure was nontoxic and environmentally friendly materials [113]. Linwei
Ji et al. studied ternary Bi2WO6/TiO2/rGO composites for photocatalytic disinfec-
tion of E. coli bacteria. Compared to pristine TiO2 and Bi2MoO6, the synthesized
ternary compound has a much smaller bandgap, high light harvesting capacity, and
less recombination rate, which enables disruption of 93% E. coli bacteria. From
radical trapping, the experiment confirmed that the holes were responsible for the
effective degradation of E. coli bacteria. The ternary nanocomposite-based photo-
catalyst emerged as an effective material for antibacterial disinfection [114].
158 M. Shetty et al.

6.8  Summary

In this chapter, Bi-based Aurivillius compounds such as Bi2MO6 (M = Cr, Mo, W),
Bi4Ti3O12, and (BiO)2CO3 were discussed with an effort to understand the basic
crystal structure and properties. The photocatalytic mechanism of pollutant degra-
dation was discussed in detail. We reviewed the most widely used and reliable
approaches for the synthesis of Aurivillius compounds, namely, hydrothermal
approach, sol-gel method, and coprecipitation method. The effects of morphology
and surface modification of Bi-based Aurivillius compounds on photocatalytic
applications were explored. Various applications of Bi-based Aurivillius compounds
such as organic dye degradation, pharmaceutical waste degradation, and bacterial
disinfection were described. The recent advances in Bi-based Aurivillius com-
pounds based on Z-scheme-type photocatalytic degradation of organic dyes were
thoroughly discussed. Some of the recent studies on organic pollutant degradation
were presented. The Bi-containing Aurivillius material with Z-scheme-based het-
erojunction shows remarkable photocatalytic efficiency compared to a p-n junction
or type II heterojunction semiconducting materials. Degradation of various pharma-
ceuticals by Bi-based Aurivillius materials was illustrated with recent studies. The
photocatalytic reduction of heavy metals by Bi-based Aurivillius compounds was
explained. Attention was given for the highly hazardous heavy metal Cr (VI) reduc-
tion. Finally, bacterial disinfection by Bi-based Aurivillius materials was explored
with recent reports.

References

1. Luo, J., Zhang, S., Sun, M., Yang, L., Luo, S., & Crittenden, J. C. (2019). A critical review
on energy conversion and environmental remediation of photocatalysts with remodeling
crystal lattice, surface, and interface. ACS Nano, 13, 9811–9840. https://doi.org/10.1021/
acsnano.9b03649.
2. Abe, R. (2010). Recent progress on photocatalytic and photoelectrochemical water split-
ting under visible light irradiation. Journal of Photochemistry and Photobiology C:
Photochemistry Review, 11, 179–209. https://doi.org/10.1016/j.jphotochemrev.2011.02.003.
3. Khaki, M.  R. D., Shafeeyan, M.  S., Raman, A.  A. A., & Daud, W.  M. A.  W. (2017).
Application of doped photocatalysts for organic pollutant degradation - A review. Journal of
Environmental Management, 198, 78–94. https://doi.org/10.1016/j.jenvman.2017.04.099.
4. Nahar, S., Zain, M. F. M., Kadhum, A. A. H., Hasan, H. A., & Hasan, M. R. (2017). Advances
in photocatalytic CO2 reduction with water: A review. Materials (Basel), 10. https://doi.
org/10.3390/ma10060629.
5. Teets, T.  S., & Nocera, D.  G. (2011). Photocatalytic hydrogen production. Chemical
Communications, 47, 9268–9274. https://doi.org/10.1039/c1cc12390d.
6. Li, R. (2018). Photocatalytic nitrogen fixation: An attractive approach for artificial photocatal-
ysis, Cuihua Xuebao/Chinese. Journal of Catalysis, 39, 1180–1188. https://doi.org/10.1016/
S1872-­2067(18)63104-­3.
7. Ahmed, S. N., & Haider, W. (2018). Heterogeneous photocatalysis and its potential appli-
cations in water and wastewater treatment: A review. Nanotechnology, 29. https://doi.
org/10.1088/1361-­6528/aac6ea.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 159

8. Frei, H. (2017). Photocatalytic fuel production. Current Opinion in Electrochemistry, 2,


128–135. https://doi.org/10.1016/j.coelec.2017.03.009.
9. Karthikeyan, C., Arunachalam, P., Ramachandran, K., Al-Mayouf, A. M., & Karuppuchamy,
S. (2020). Recent advances in semiconductor metal oxides with enhanced methods for solar
photocatalytic applications. Journal of Alloys and Compounds, 828, 154281. https://doi.
org/10.1016/j.jallcom.2020.154281.
10. Shamraiz, U., Hussain, R. A., Badshah, A., Raza, B., & Saba, S. (2016). Functional metal sul-
fides and selenides for the removal of hazardous dyes from water. Journal of Photochemistry
and Photobiology B: Biology, 159, 33–41. https://doi.org/10.1016/j.jphotobiol.2016.03.013.
11. Cao, S., Wang, C. J., Fu, W. F., & Chen, Y. (2017). Metal phosphides as co-catalysts for pho-
tocatalytic and photoelectrocatalytic water splitting. ChemSusChem, 10, 4306–4323. https://
doi.org/10.1002/cssc.201701450.
12. Liang, J., Chen, D., Yao, X., Zhang, K., Qu, F., Qin, L., Huang, Y., & Li, J. (2020). Recent
progress and development in inorganic halide Perovskite quantum dots for photoelectro-
chemical applications. Small, 16, 1–20. https://doi.org/10.1002/smll.201903398.
13. Sharma, K., Dutta, V., Sharma, S., Raizada, P., Hosseini-Bandegharaei, A., Thakur, P., &
Singh, P. (2019). Recent advances in enhanced photocatalytic activity of bismuth oxyhalides
for efficient photocatalysis of organic pollutants in water: A review. Journal of Industrial and
Engineering Chemistry, 78, 1–20. https://doi.org/10.1016/j.jiec.2019.06.022.
14. Zhou, W., Yan, L., Wang, Y., & Zhang, Y. (2006). SiC nanowires: A photocatalytic nanomate-
rial. Applied Physics Letters, 89. https://doi.org/10.1063/1.2219139.
15. Dong, G., Zhang, Y., Pan, Q., & Qiu, J. (2014). A fantastic graphitic carbon nitride (g-C3N4)
material: Electronic structure, photocatalytic and photoelectronic properties. Journal
of Photochemistry and Photobiology C: Photochemistry Review, 20, 33–50. https://doi.
org/10.1016/j.jphotochemrev.2014.04.002.
16. Liu, D., Ma, J., Long, R., Gao, C., & Xiong, Y. (2017). Silicon nanostructures for solar-driven
catalytic applications. Nano Today, 17, 96–116. https://doi.org/10.1016/j.nantod.2017.10.013.
17. Ong, C. B., Ng, L. Y., & Mohammad, A. W. (2018). A review of ZnO nanoparticles as solar
photocatalysts: Synthesis, mechanisms and applications. Renewable and Sustainable Energy
Reviews, 81, 536–551. https://doi.org/10.1016/j.rser.2017.08.020.
18. Qi, K., Cheng, B., Yu, J., & Ho, W. (2017). A review on TiO2-based Z-scheme photocata-
lysts. Cuihua Xuebao/Chinese Journal of Catalysis, 38, 1936–1955. https://doi.org/10.1016/
S1872-­2067(17)62962-­0.
19. Malathi, A., Madhavan, J., Ashokkumar, M., & Arunachalam, P. (2018). A review on BiVO4
photocatalyst: Activity enhancement methods for solar photocatalytic applications. Applied
Catalysis A: General, 555, 47–74. https://doi.org/10.1016/j.apcata.2018.02.010.
20. Tang, Z. R., Han, B., Han, C., & Xu, Y. J. (2017). One dimensional CdS based materials for
artificial photoredox reactions. Journal of Materials Chemistry A, 5, 2387–2410. https://doi.
org/10.1039/C6TA06373J.
21. Liu, X., Gu, S., Zhao, Y., Zhou, G., & Li, W. (2020). BiVO4, Bi2WO6 and Bi2MoO6 photocatal-
ysis: A brief review. Journal of Materials Science and Technology. https://doi.org/10.1016/j.
jmst.2020.04.023.
22. Pan, L., Zou, J. J., Zhang, T., Wang, S., Li, Z., Wang, L., & Zhang, X. (2014). Cu2O film
via hydrothermal redox approach: Morphology and photocatalytic performance. Journal of
Physical Chemistry C, 118, 16335–16343. https://doi.org/10.1021/jp408056k.
23. He, R., Xu, D., Cheng, B., Yu, J., & Ho, W. (2018). Review on nanoscale Bi-based photocata-
lysts. Nanoscale Horizons, 3, 464–504. https://doi.org/10.1039/c8nh00062j.
24. Zhou, L., Wang, W., Xu, H., Sun, S., & Shang, M. (2009). Bi2O3 hierarchical nanostruc-
tures: Controllable synthesis, growth mechanism, and their application in photocatalysis.
Chemistry – A European Journal, 15, 1776–1782. https://doi.org/10.1002/chem.200801234.
25. Wang, L., Ma, W., Fang, Y., Zhang, Y., Jia, M., Li, R., & Huang, Y. (2013). Bi4Ti3O12
Synthesized by high temperature solid phase method and it’s visible catalytic activity. Procedia
Environmental Sciences, 18, 547–558. https://doi.org/10.1016/j.proenv.2013.04.074.
160 M. Shetty et al.

26. Huang, Y., Li, K., Lin, Y., Tong, Y., & Liu, H. (2018). Enhanced efficiency of electron–hole
separation in Bi2O2CO3 for photocatalysis via acid treatment. ChemCatChem, 10, 1982–1987.
https://doi.org/10.1002/cctc.201800101.
27. Dong, S., Ding, X., Guo, T., Yue, X., Han, X., & Sun, J. (2017). Self-assembled hollow sphere
shaped Bi2WO6/RGO composites for efficient sunlight-driven photocatalytic degradation of
organic pollutants. Chemical Engineering Journal, 316, 778–789. https://doi.org/10.1016/j.
cej.2017.02.017.
28. Zhou, L., Wang, W., & Zhang, L. (2007). Ultrasonic-assisted synthesis of visible-light-­
induced Bi2MO6 (M = W, Mo) photocatalysts. Journal of Molecular Catalysis A: Chemical,
268, 195–200. https://doi.org/10.1016/j.molcata.2006.12.026.
29. Lv, Y., Liu, Y., Zhu, Y., & Zhu, Y. (2014). Surface oxygen vacancy induced photocatalytic per-
formance enhancement of a BiPO4 nanorod. Journal of Materials Chemistry A, 2, 1174–1182.
https://doi.org/10.1039/c3ta13841k.
30. Gao, F., Chen, X., Yin, K., Dong, S., Ren, Z., Yuan, F., Yu, T., Zou, Z., & Liu, J. M. (2007).
Visible-light photocatalytic properties of weak magnetic BiFeO3 nanoparticles. Advanced
Materials, 19, 2889–2892. https://doi.org/10.1002/adma.200602377.
31. Yin, H., Zhou, A., Chang, N., & Xu, X. (2009). Characterization and photocatalytic activity
of Bi3TiNbO9 nanocrystallines synthesized by sol–gel process. Materials Research Bulletin,
44, 377–380. https://doi.org/10.1016/j.materresbull.2008.05.008.
32. Bac, L. H., Thanh, L. T. H., Van Chinh, N., Khoa, N. T., Van Thiet, D., Van Trung, T., &
Dung, D. D. (2016). Tailoring the structural, optical properties and photocatalytic behavior
of ferroelectric Bi0.5K0.5TiO3 nanopowders. Materials Letters, 164, 631–635. https://doi.
org/10.1016/j.matlet.2015.11.086.
33. Ramachandran, R., Sathiya, M., Ramesha, K., Prakash, A.  S., Madras, G., & Shukla,
A. K. (2011). Photocatalytic properties of KBiO3 and LiBiO3 with tunnel structures. Journal
of Chemical Sciences, 123, 517–524. https://doi.org/10.1007/s12039-­011-­0080-­9.
34. Yu, K., Yang, S., He, H., Sun, C., Gu, C., & Ju, Y. (2009). Visible light-driven photocata-
lytic degradation of rhodamine B over NaBiO3: Pathways and mechanism. The Journal of
Physical Chemistry. A, 113, 10024–10032. https://doi.org/10.1021/jp905173e.
35. Hincapié, C.  M. B., Cárdenas, M.  J. P., Orjuela, J.  E. A., Parra, E.  R., & Olaya Florez,
J. J. (2012). Physical-chemical properties of bismuth and bismuth oxides: Synthesis, charac-
terization and applications | Propiedades físico-químicas del bismuto y oxidos de bismuto:
Síntesis, caracterización y aplicaciones. DYNA, 79, 139–148.
36. Pirovano, C., Islam, M.  S., Vannier, R.  N., Nowogrocki, G., & Mairesse, G. (2001).
Mophasesdelling the crystal structures of Aurivillius. Solid State Ionics, 140, 115–123.
https://doi.org/10.1016/S0167-­2738(01)00699-­3.
37. Li, Z., Zhang, Z., Wang, L., & Meng, X. (2020). Bismuth chromate (Bi2CrO6): A promising
semiconductor in photocatalysis. Journal of Catalysis, 382, 40–48. https://doi.org/10.1016/j.
jcat.2019.12.001.
38. Meng, X., & Zhang, Z. (2016). Bismuth-based photocatalytic semiconductors: Introduction,
challenges and possible approaches. Journal of Molecular Catalysis A: Chemical, 423,
533–549. https://doi.org/10.1016/j.molcata.2016.07.030.
39. Singh, D.  J., Seo, S.  S. A., & Lee, H.  N. (2010). Optical properties of ferroelectric Bi4
Ti3O12. Physical Review B: Condensed Matter and Materials Physics, 82, 3–6. https://doi.
org/10.1103/PhysRevB.82.180103.
40. Chen, X., Shen, S., Guo, L., & Mao, S. S. (2010). Semiconductor-based photocatalytic hydro-
gen generation. Chemical Reviews, 110, 6503–6570. https://doi.org/10.1021/cr1001645.
41. Cui, M. Y., He, J. X., Lu, N. P., Zheng, Y. Y., Dong, W. J., Tang, W. H., Chen, B. Y., & Li,
C. R. (2010). Morphology and size control of cerium carbonate hydroxide and ceria micro/
nanostructures by hydrothermal technology. Materials Chemistry and Physics, 121, 314–319.
https://doi.org/10.1016/j.matchemphys.2010.01.041.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 161

42. H.S. Neira, Ines S., Kolen’ko, Yury V., Lebedev, Oleg I., Tendeloo, Gustaaf Van, Gupta, F. &
Y.M.  Guitian (2009) . An effective morphology control of hydroxyapatite crystals via &
design ND, An Eff. Morphol. Control Hydroxyapatite Cryst. via Hydrothermal Synth.
43. Liu, Z., Sun, D. D., Guo, P., & Leckie, J. O. (2007). One-step fabrication and high photo-
catalytic activity of porous TiO2 hollow aggregates by using a low-temperature hydrothermal
method without templates. Chemistry  – A European Journal, 13, 1851–1855. https://doi.
org/10.1002/chem.200601092.
44. Neto, N. F. A., Nunes, T. B. O., Li, M., Longo, E., Bomio, M. R. D., & Motta, F. V. (2020).
Influence of microwave-assisted hydrothermal treatment time on the crystallinity, mor-
phology and optical properties of ZnWO4 nanoparticles: Photocatalytic activity. Ceramics
International, 46, 1766–1774. https://doi.org/10.1016/j.ceramint.2019.09.151.
45. Hirano, M., & Okamoto, T. (2018). Synthesis, morphology, and luminescence of ZnNb2O6
nanocrystals by hydrothermal method. Nano-Structures and Nano-Objects, 13, 139–145.
https://doi.org/10.1016/j.nanoso.2015.11.001.
46. Fabrication, S., & Feo, A. (2006). Crystal growth & design 2006. Communications, 6, 2–4.
47. Bai, J., Li, Y., Li, X., & Liu, L. (2017). Facile preparation of 2D Bi2MoO6 nanosheets-RGO
composites with enhanced photocatalytic activity. New Journal of Chemistry, 41, 7783–7790.
https://doi.org/10.1039/c7nj01712j.
48. Cai, K., Lv, S. Y., Song, L. N., Chen, L., He, J., Chen, P., Au, C. T., & Yin, S. F. (2019).
Facile preparation of ultrathin Bi2MoO6 nanosheets for photocatalytic oxidation of toluene to
benzaldehyde under visible light irradiation. Journal of Solid State Chemistry, 269, 145–150.
https://doi.org/10.1016/j.jssc.2018.09.027.
49. Silva Ribeiro, C., & Azário Lansarin, M. (2019). Facile solvo-hydrothermal synthesis of
Bi2MoO6 for the photocatalytic reduction of CO2 into ethanol in water under visible light.
Reaction Kinetics, Mechanisms and Catalysis, 127, 1059–1071. https://doi.org/10.1007/
s11144-­019-­01591-­z.
50. Zhao, Y., Wang, Y., Liu, E., Fan, J., & Hu, X. (2018). Bi2WO6 nanoflowers: An efficient visi-
ble light photocatalytic activity for ceftriaxone sodium degradation. Applied Surface Science,
436, 854–864. https://doi.org/10.1016/j.apsusc.2017.12.064.
51. Yang, A.  M., Han, Y., Li, S.  S., Xing, H.  W., Pan, Y.  H., & Liu, W.  X. (2017). Synthesis
and comparison of photocatalytic properties for Bi2WO6 nanofibers and hierarchical
microspheres. Journal of Alloys and Compounds, 695, 915–921. https://doi.org/10.1016/j.
jallcom.2016.10.188.
52. Tseng, T. K., Lin, Y. S., Chen, Y. J., & Chu, H. (2010). A review of photocatalysts prepared
by sol-gel method for VOCs removal. International Journal of Molecular Sciences, 11,
2336–2361. https://doi.org/10.3390/ijms11062336.
53. Liu, Y., Lv, H., Hu, J., & Li, Z. (2015). Synthesis and characterization of Bi2WO6 nanoplates
using egg white as a biotemplate through sol–gel method. Materials Letters, 139, 401–404.
https://doi.org/10.1016/j.matlet.2014.10.131.
54. Guo, Y., Zhang, G., & Gan, H. (2012). Synthesis, characterization and visible light pho-
tocatalytic properties of Bi2WO6/rectorite composites. Journal of Colloid and Interface
Science, 369, 323–329. https://doi.org/10.1016/j.jcis.2011.11.066.
55. Waqar, M., Imran, M., Adil, S. F., Noreen, S., Latif, S., Khan, M., & Siddiqui, M. R. H. (2020).
Enhanced photoluminescence and photocatalytic efficiency of la-doped bismuth molyb-
date: Its preparation and characterization. Materials (Basel), 13. https://doi.org/10.3390/
ma13010035.
56. Fuentes, S., Muñoz, P., Llanos, J., Vega, M., Martin, I.  R., & Chavez-Angel, E. (2017).
Synthesis and optical characterization of Er-doped bismuth titanate nanoparticles grown
by sol–gel hydrothermal method. Ceramics International, 43, 3623–3630. https://doi.
org/10.1016/j.ceramint.2016.11.200.
57. Li, L., Ma, Z., Bi, F., Li, H., Zhang, W., & He, H. (2016). Sol–gel preparation and prop-
erties of Bi4Ti3O12 photocatalyst supported on micrometer-sized quartz spheres. Journal of
Advanced Oxidation Technologies, 19, 310–316.
162 M. Shetty et al.

58. Luévano-Hipólito, E., de la Cruz, A. M., & Cuéllar, E. L. (2014). Synthesis, characterization,
and photocatalytic properties of γ-Bi2MoO6 prepared by co-precipitation assisted with ultra-
sound irradiation. Journal of the Taiwan Institute of Chemical Engineers, 45, 2749–2754.
https://doi.org/10.1016/j.jtice.2014.05.024.
59. Banavatu, L., Rao, D. S., & Basavaiah, K. (2018). Synthesis of γ-Bi2MoO6 by co-­precipitation
method and evaluation for photocatalytic degradation of rhodamine B, crystal violet and
orange II dyes under visible light irradiation. Asian Journal of Chemistry, 30, 97–102. https://
doi.org/10.14233/ajchem.2018.20917.
60. Ait Ahsaine, H., Ezahri, M., Benlhachemi, A., Bakiz, B., Villain, S., Guinneton, F., &
Gavarri, J. R. (2016). Novel Lu-doped Bi2WO6 nanosheets: Synthesis, growth mechanisms
and enhanced photocatalytic activity under UV-light irradiation. Ceramics International, 42,
8552–8558. https://doi.org/10.1016/j.ceramint.2016.02.082.
61. Li, S., Zhang, J., Chen, X., Feng, Z., & Li, C. (2017). Effect of surface loading and bulk
doping of La3+ on the thermal stability and photocatalytic activity of Bi2O2CO3. Materials
Research Bulletin, 94, 127–133. https://doi.org/10.1016/j.materresbull.2017.05.058.
62. Guo, C., Xu, J., Wang, S., Li, L., Zhang, Y., & Li, X. (2012). Facile synthesis and photo-
catalytic application of hierarchical mesoporous Bi2MoO6 nanosheet-based microspheres.
CrystEngComm, 14, 3602–3608. https://doi.org/10.1039/c2ce06757a.
63. Zhang, L., Wang, H., Chen, Z., Wong, P. K., & Liu, J. (2011). Bi2WO6 micro/nano-­structures:
Synthesis, modifications and visible-light-driven photocatalytic applications. Applied
Catalysis B: Environmental, 106, 1–13. https://doi.org/10.1016/j.apcatb.2011.05.008.
64. Zhang, L., Wang, H., Chen, Z., Keung, P., & Liu, J. (2011). Environmental Bi2WO6 micro/
nano-structures: Synthesis, modifications and visible-light-driven photocatalytic applications.
Applied Catalysis B: Environmental, 106, 1–13. https://doi.org/10.1016/j.apcatb.2011.05.008.
65. Zhao, T., Zai, J., Xu, M., Zou, Q., Su, Y., Wang, K., & Qian, X. (2011). Hierarchical Bi2O2CO3
microspheres with improved visible-light-driven photocatalytic activity. CrystEngComm, 13,
4010–4017. https://doi.org/10.1039/c1ce05113j.
66. Jing, K., Ma, W., Ren, Y., Xiong, J., Guo, B., Song, Y., Liang, S., & Wu, L. (2019). Hierarchical
Bi2MoO6 spheres in situ assembled by monolayer nanosheets toward photocatalytic selective
oxidation of benzyl alcohol. Applied Catalysis B: Environmental, 243, 10–18. https://doi.
org/10.1016/j.apcatb.2018.10.027.
67. Mu, J. J., Zheng, G. H., & Ma, Y. Q. (2017). Morphology and photocatalytic properties of
γ-Bi2MoO6 tuned by stirring and surfactant EDTA assistant. Journal of Electronic Materials,
46, 596–601. https://doi.org/10.1007/s11664-­016-­4925-­3.
68. Xu, C., Zou, D., Wang, L., Luo, H., & Ying, T. (2009). γ-Bi2MoO6 nanoplates: Surfactant-­
assisted hydrothermal synthesis and optical properties. Ceramics International, 35,
2099–2102. https://doi.org/10.1016/j.ceramint.2008.11.016.
69. Zhang, P., Hua, X., Teng, X., Liu, D., Qin, Z., & Ding, S. (2016). CTAB assisted hydro-
thermal synthesis of lamellar Bi2WO6 with superior photocatalytic activity for rhodamine b
degradation. Materials Letters, 185, 275–277. https://doi.org/10.1016/j.matlet.2016.08.148.
70. Wei, S. Y., Li, M. K., Shang, D., Wang, N., Sun, L. X., & Xing, Y. H. (2015). 3D Hierarchical
flower-like Bi2WO6 microparticles doped with uranyl group: Characterization and higher
photocatalytic activity. Journal of Inorganic and Organometallic Polymers and Materials,
25, 1434–1440. https://doi.org/10.1007/s10904-­015-­0256-­8.
71. Yuanyuan Li, G. L., Liu, J., & Huang, X. (2007). Hydrothermal synthesis of Bi2WO6 uni-
form hierarchical microspheres. Crystal Growth & Design, 07, 1350–1355. https://doi.
org/10.1016/j.matlet.2008.10.012.
72. Zhang, G.  Y., Wang, J.  J., Shen, X.  Q., Wang, J.  J., Wang, B.  Y., & Gao, D.  Z. (2019).
Br-doped Bi2O2 CO3 nanosheets with improved electronic structure and accelerated charge
migration for outstanding photocatalytic behavior. Applied Surface Science, 470, 63–73.
https://doi.org/10.1016/j.apsusc.2018.11.103.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 163

73. Anku, W.  W., Oppong, S.  O. B., & Govender, P.  P. (2018). Bismuth-based nanoparticles
as photocatalytic materials, Bismuth – Advanced applications and defects characterization.
IntechOpen. https://doi.org/10.5772/intechopen.75104.
74. Zhang, L., Tan, G., Wei, S., Ren, H., Xia, A., & Luo, Y. (2013). Microwave hydrothermal
synthesis and photocatalytic properties of TiO2/BiVO4 composite photocatalysts. Ceramics
International, 39, 8597–8604. https://doi.org/10.1016/j.ceramint.2013.03.106.
75. Cho, M., Chung, H., Choi, W., & Yoon, J. (2004). Linear correlation between inactivation of
E. coli and OH radical concentration in TiO2 photocatalytic disinfection. Water Research, 38,
1069–1077. https://doi.org/10.1016/j.watres.2003.10.029.
76. Mosleh, S., Dashtian, K., Ghaedi, M., & Amiri, M. (2019). A Bi2WO6/Ag2S/ZnS: Z-scheme
heterojunction photocatalyst with enhanced visible-light photoactivity towards the degrada-
tion of multiple dye pollutants. RSC Advances, 9, 30100–30111. https://doi.org/10.1039/
c9ra05372g.
77. Xia, K., Chen, H., Mao, M., Chen, Z., Xu, F., Yi, J., Yu, Y., She, X., Xu, H., & Li, H. (2018).
Designing visible-light-driven Z-scheme catalyst 2D g-C3N4/Bi2MoO6: Enhanced photodeg-
radation activity of organic pollutants. Physica Status Solidi (A) Applications and Materials
Science, 215, 2–9. https://doi.org/10.1002/pssa.201800520.
78. Jia, J., Du, X., Zhang, Q., Liu, E., & Fan, J. (2019). Z-scheme MgFe2O4/Bi2MoO6 het-
erojunction photocatalyst with enhanced visible light photocatalytic activity for mala-
chite green removal. Applied Surface Science, 492, 527–539. https://doi.org/10.1016/j.
apsusc.2019.06.258.
79. Liu, H., Du, C., Bai, H., Su, Y., Wei, D., Wang, Y., Liu, G., & Yang, L. (2018). Fabrication of
plate-on-plate Z-scheme SnS2/Bi2MoO6 heterojunction photocatalysts with enhanced photo-
catalytic activity. Journal of Materials Science, 53, 10743–10757. https://doi.org/10.1007/
s10853-­018-­2296-­2.
80. Li, Z., Meng, X., & Zhang, Z. (2018). Few-layer MoS2 nanosheets-deposited on Bi2MoO6
microspheres: A Z-scheme visible-light photocatalyst with enhanced activity. Catalysis
Today, 315, 67–78. https://doi.org/10.1016/j.cattod.2018.03.014.
81. Ji, M., Di, J., Liu, Y., Chen, R., Li, K., Chen, Z., Xia, J., & Li, H. (2020). Confined active
species and effective charge separation in Bi4O5I2 ultrathin hollow nanotube with increased
photocatalytic activity. Applied Catalysis B: Environmental, 268, 118403. https://doi.
org/10.1016/j.apcatb.2019.118403.
82. Tao, R., Shao, C., Li, X., Li, X., Liu, S., Yang, S., Zhao, C., & Liu, Y. (2018). Bi2MoO6/
BiFeO3 heterojunction nanofibers: Enhanced photocatalytic activity, charge separation mech-
anism and magnetic separability. Journal of Colloid and Interface Science, 529, 404–414.
https://doi.org/10.1016/j.jcis.2018.06.035.
83. Zhang, J., Niu, C., Ke, J., Zhou, L., & Zeng, G. (2015). Ag/AgCl/Bi2MoO6 composite
nanosheets: A plasmonic Z-scheme visible light photocatalyst. Catalysis Communications,
59, 30–34. https://doi.org/10.1016/j.catcom.2014.09.041.
84. Wang, D., Shen, H., Guo, L., Fu, F., & Liang, Y. (2016). Design and construction of the
sandwich-like Z-scheme multicomponent CdS/Ag/Bi2MoO6 heterostructure with enhanced
photocatalytic performance in RhB photodegradation. New Journal of Chemistry, 40,
8614–8624. https://doi.org/10.1039/c6nj01893a.
85. Kasinathan, M., Thiripuranthagan, S., Sivakumar, A., Ranganathan, S., Vembuli, T.,
Kumaravel, S., & Erusappan, E. (2020). Fabrication of novel Bi2MoO6/N-rGO catalyst for
the efficient photocatalytic degradation of harmful dyes. Materials Research Bulletin, 125,
110782. https://doi.org/10.1016/j.materresbull.2020.110782.
86. Liu, Y., Yang, Z. H., Song, P. P., Xu, R., & Wang, H. (2018). Facile synthesis of Bi2MoO6/
ZnSnO3 heterojunction with enhanced visible light photocatalytic degradation of methylene
blue. Applied Surface Science, 430, 561–570. https://doi.org/10.1016/j.apsusc.2017.06.231.
87. Chou, X., Ye, J., He, J., Ge, K., Liu, J., Fu, C., Zhou, X., Wang, S., Zhang, Y., & Yang,
Y. (2019). One-step solvothermal synthesis of BiPO4/Bi2MoO6 heterostructure with oxygen
164 M. Shetty et al.

vacancies and Z-Scheme system for enhanced photocatalytic performance. ChemistrySelect,


4, 8327–8333. https://doi.org/10.1002/slct.201901433.
88. Wang, Z., Lv, J., Dai, K., Lu, L., Liang, C., & Geng, L. (2016). Large scale and facile synthe-
sis of novel Z-scheme Bi2MoO6/Ag3PO4 composite for enhanced visible light photocatalyst.
Materials Letters, 169, 250–253. https://doi.org/10.1016/j.matlet.2016.01.147.
89. Chankhanittha, T., Somaudon, V., Watcharakitti, J., Piyavarakorn, V., & Nanan, S. (2020).
Performance of solvothermally grown Bi2MoO6 photocatalyst toward degradation of organic
azo dyes and fluoroquinolone antibiotics. Materials Letters, 258, 126764. https://doi.
org/10.1016/j.matlet.2019.126764.
90. Tian, J., Zhu, Z., & Liu, B. (2019). Novel Bi2MoO6/Bi2WO6/MWCNTs photocatalyst with
enhanced photocatalytic activity towards degradation of RB-19 under visible light irradia-
tion. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 581, 123798.
https://doi.org/10.1016/j.colsurfa.2019.123798.
91. Xu, L. J., & Wang, J. L. (2013). file:///F:/book chapter Bi/organic chemicals/wang2016Syner-
gistic photocatalysis o.pdf. Applied Catalysis B: Environmental, 142–143, 396–405. https://
doi.org/10.1016/j.apcatb.2013.05.065.
92. Wang, J.  C., Ren, J., Yao, H.  C., Zhang, L., Wang, J.  S., Zang, S.  Q., Han, L.  F., & Li,
Z. J. (2016). Synergistic photocatalysis of Cr(VI) reduction and 4-Chlorophenol degradation
over hydroxylated α-Fe2O3 under visible light irradiation. Journal of Hazardous Materials,
311, 11–19. https://doi.org/10.1016/j.jhazmat.2016.02.055.
93. Ma, T., Wu, J., Mi, Y., Chen, Q., Ma, D., & Chai, C. (2017). Novel Z-Scheme g-C3N4/C@
Bi2MoO6 composite with enhanced visible-light photocatalytic activity for Β-naphthol deg-
radation. Separation and Purification Technology, 183, 54–65. https://doi.org/10.1016/j.
seppur.2017.04.005.
94. Cui, Y., Zhang, H., Guo, R., Ma, Q., Deng, X., Cheng, X., Li, X., Xie, M., Cheng, Q., & Zou,
C. (2017). Fabrication of bismuth oxide/titanium dioxide nano-tube arrays photoelectrode
and its enhanced visible light photocatalytic performance for degradation of 4-chlorphenol.
Electrochimica Acta, 246, 1075–1081. https://doi.org/10.1016/j.electacta.2017.06.153.
95. Wu, X., Zhang, Y., Wang, K., Zhang, S., Qu, X., Shi, L., & Du, F. (2020). In-situ construc-
tion of Bi/defective Bi4NbO8Cl for non-noble metal based Mott-Schottky photocatalysts
towards organic pollutants removal. Journal of Hazardous Materials, 393, 122408. https://
doi.org/10.1016/j.jhazmat.2020.122408.
96. Yang, Q., Guo, E., Liu, H., & Lu, Q. (2019). Engineering of Z-scheme 2D/3D architectures
with Bi2MoO6 on TiO2 nanosphere for enhanced photocatalytic 4-nitrophenol degradation.
Journal of the Taiwan Institute of Chemical Engineers, 105, 65–74. https://doi.org/10.1016/j.
jtice.2019.09.024.
97. Fu, F., Shen, H., Xue, W., Zhen, Y., Soomro, R.  A., Yang, X., Wang, D., Xu, B., & Chi,
R. (2019). Alkali-assisted synthesis of direct Z-scheme based Bi2O3/Bi2MoO6 photocatalyst
for highly efficient photocatalytic degradation of phenol and hydrogen evolution reaction.
Journal of Catalysis, 375, 399–409. https://doi.org/10.1016/j.jcat.2019.06.033.
98. Rong, Q., Zhang, D., Li, Y., Zha, Z., Geng, X., Cui, S., & Yang, J. (2019). Synthesis of
Bi2MoO6/Bi2Ti2O7 Z-Scheme heterojunction as efficient visible-light photocatalyst for the
glycolic acid degradation. Journal of Nanoscience and Nanotechnology, 19, 7635–7644.
https://doi.org/10.1166/jnn.2019.16779.
99. Shi, B., Yin, H., Gong, J., & Nie, Q. (2017). Ag/AgCl decorated Bi4Ti3O12 nanosheet with
highly exposed (001) facets for enhanced photocatalytic degradation of Rhodamine B, carba-
mazepine and tetracycline. Applied Surface Science, 419, 614–623. https://doi.org/10.1016/j.
apsusc.2017.05.103.
100. Xue, W., Peng, Z., Huang, D., Zeng, G., Wen, X., Deng, R., Yang, Y., & Yan, X. (2019).
In situ synthesis of visible-light-driven Z-scheme AgI/Bi2WO6 heterojunction photocatalysts
with enhanced photocatalytic activity. Ceramics International, 45, 6340–6349. https://doi.
org/10.1016/j.ceramint.2018.12.119.
6  Photocatalytic Efficiency of Bi-Based Aurivillius Compounds: Critical Review… 165

101. Liang, W., Pan, J., Duan, X., Tang, H., Xu, J., & Tang, G. (2020). Biomass carbon modi-
fied flower-like Bi2WO6 hierarchical architecture with improved photocatalytic performance.
Ceramics International, 46, 3623–3630. https://doi.org/10.1016/j.ceramint.2019.10.081.
102. Liu, X., Zhou, Z., Han, D., Wang, T., Ma, C., Huo, P., & Yan, Y. (2020). Interface engineered
2D/2D Ni(OH)2/Bi4Ti3O12 nanocomposites with higher charge transfer towards improv-
ing photocatalytic activity. Journal of Alloys and Compounds, 816, 152530. https://doi.
org/10.1016/j.jallcom.2019.152530.
103. Li, H., Zhang, Y., Ou, H., Ma, T., & Huang, H. (2020). Two layered Bi-based borate pho-
tocatalysts MBi2B2O7 (M  =  Ca, Sr) for photocatalytic degradation and oxygen activation.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 584, 123994. https://
doi.org/10.1016/j.colsurfa.2019.123994.
104. Liu, Y., Liu, S., Wu, T., Lin, H., & Zhang, X. (2017). Facile preparation of flower-like
Bi2WO6/CdS heterostructured photocatalyst with enhanced visible-light-driven photocata-
lytic activity for Cr(VI) reduction. Journal of Sol–Gel Science and Technology, 83, 315–323.
https://doi.org/10.1007/s10971-­017-­4416-­x.
105. Yang, J., Wang, X., Chen, Y., Dai, J., & Sun, S. (2015). Enhanced photocatalytic activities of
visible-light driven green synthesis in water and environmental remediation on Au/Bi2WO6
hybrid nanostructures. RSC Advances, 5, 9771–9782. https://doi.org/10.1039/c4ra15349a.
106. Rauf, A., Sher Shah, M.  S. A., Choi, G.  H., Bin Humayoun, U., Yoon, D.  H., Bae, J.  W.,
Park, J., Kim, W. J., & Yoo, P. J. (2015). Facile synthesis of hierarchically structured Bi2S3/
Bi2WO6 photocatalysts for highly efficient reduction of Cr(VI). ACS Sustainable Chemistry
& Engineering, 3, 2847–2855. https://doi.org/10.1021/acssuschemeng.5b00783.
107. Ri, C. N., Song-Gol, K., Ju-Yong, J., Pak, S. N., Ri, S. C., & Ri, J. H. (2018). Construction
of the Bi2WO6/Bi4V2O11 heterojunction for highly efficient visible-light-driven photocata-
lytic reduction of Cr(VI). New Journal of Chemistry, 42, 647–653. https://doi.org/10.1039/
c7nj03413j.
108. Meng, Q., Zhou, Y., Chen, G., Hu, Y., Lv, C., Qiang, L., & Xing, W. (2018). Integrating
both homojunction and heterojunction in QDs self-decorated Bi2MoO6/BCN composites to
achieve an efficient photocatalyst for Cr(VI) reduction. Chemical Engineering Journal, 334,
334–343. https://doi.org/10.1016/j.cej.2017.07.134.
109. Yang, J., Wang, X., Zhao, X., Dai, J., & Mo, S. (2015). Synthesis of uniform Bi2WO6-reduced
graphene oxide nanocomposites with significantly enhanced photocatalytic reduction activ-
ity. Journal of Physical Chemistry C, 119, 3068–3078. https://doi.org/10.1021/jp510041x.
110. Pulgarin, C., Karbasi, M., Karimzadeh, F., Raeissi, K., Rtimi, S., Giannakis, S., & Kiwi,
J. (2020). Insights into the photocatalytic bacterial inactivation by flower-like Bi2WO6 under
solar or visible light. Water, 12, 1–19. https://doi.org/10.3390/w12041099.
111. Li, M., Li, D., Zhou, Z., Wang, P., Mi, X., Xia, Y., Wang, H., Zhan, S., Li, Y., & Li, L. (2020).
Plasmonic Ag as electron-transfer mediators in Bi2MoO6/Ag–AgCl for efficient photocata-
lytic inactivation of bacteria. Chemical Engineering Journal, 382. https://doi.org/10.1016/j.
cej.2019.122762.
112. Shanmugam, V., Muppudathi, A. L., Jayavel, S., & Jeyaperumal, K. S. (2020). Construction
of high efficient g-C3N4 nanosheets combined with Bi2MoO6-Ag photocatalysts for visible-­
light-­driven photocatalytic activity and inactivation of bacterias. Arabian Journal of
Chemistry, 13, 2439–2455. https://doi.org/10.1016/j.arabjc.2018.05.009.
113. Shi, H., Fan, J., Zhao, Y., Hu, X., Zhang, X., & Tang, Z. (2020). Visible light driven CuBi2O4/
Bi2MoO6 p-n heterojunction with enhanced photocatalytic inactivation of E. coli and mech-
anism insight. Journal of Hazardous Materials, 381, 121006. https://doi.org/10.1016/j.
jhazmat.2019.121006.
114. Ji, L., Liu, B., Qian, Y., Yang, Q., & Gao, P. (2020). Enhanced visible-light-induced photo-
catalytic disinfection of Escherichia coli by ternary Bi2WO6/TiO2/reduced graphene oxide
composite materials: Insight into the underlying mechanism. Advanced Powder Technology,
31, 128–138. https://doi.org/10.1016/j.apt.2019.10.005.
Chapter 7
Intrinsically Conducting Polymer
Nanocomposites in Shielding
of Electromagnetic Pollution

Suneel Kumar Srivastava

7.1  General Introduction

The interference of electromagnetic (EM) waves is an undesirable manifestation of


different emitted frequency radiation due to extensive use of electronic and wireless
electronic devices, such as mobile phones, computers, TV, radio, etc. [1–19]. The
interference of these radiations with different electronic instruments and appliances
deteriorate their performance in terms of speed and secrecy resulting in the loss of
data storage, revenue, efforts and time. In addition, it makes adverse effect on
human health resulting in many diseases, such as leukaemia, miscarriages and brain
cancer and needs immediate solution. Therefore, electromagnetic interference
shielding is essentially needed in a variety of industries including automotive, elec-
tronics, energy and healthcare. The shielding materials should be flexible, light
weight, low-cost and easily available and high performing in preventing the inter-
ference of electromagnetic radiation from other devices.
In view of this, the large number of metallic materials has been investigated as
effective EMI shielding materials [20]. Their performance is attributed to the
interaction of electron cloud near the metal surface with the striking EM wave.
These result in the distortion of electron cloud and generation of electric field in
the direction opposite to the applied electric field. This causes reflection of EM
wave from the surface rather than penetration through the metal due to impedance
mismatch. However, heavy weight, corrosion, and poor processability of metals
were few important drawbacks [20, 21]. Recently, attention has been focused on
intrinsically conducting polymers (ICPs) as a substitute for conventional EMI
shielding materials due to their portability, low cost/density, conductivity control,
adjustable permittivity/permeability and corrosion resistance as well as good
thermal and environmental stability [23–27]. The common examples of ICPs

S. K. Srivastava (*)
Department of Chemistry, Indian Institute of Technology, Kharagpur,
West Bengal, India

© Springer Nature Switzerland AG 2021 167


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_7
168 S. K. Srivastava

Fig. 7.1  Structures of


common intrinsically
conducting polymers

include polyaniline (PANI), polypyrrole (PPy), polythiophene (PTh), polypropyl-


ene (PP), polyparaphenylene (PPP) and polyparaphenylenevinylene (PPV).
Figure  7.1 shows the schematic chemical structures of different the ICPs. The
ICPs display unique physical and electrical properties due to their chemical struc-
tures. In particular, PANI among the available ICPs remains one of the most
promising EMI shielding/absorbing material [26, 28]. The frequency dependence
of intrinsic conductivity in microwave region marks an added advantage in their
EMI shielding applications [28]. Investigations have shown that ICPs combined
with conducting magnetic dielectric materials further enhanced their EMI shield-
ing performance. It is inferred that high permittivity/permeability of inorganic
materials could make realization of ICP nanocomposites with superior EMI
shielding properties.
ICPs are commonly synthesized by oxidative chemical polymerization of aniline
monomer in acidic condition or by electrochemical polymerization. The available
literature suggests extensive work being reported on PANI/PPy compared to any
other ICPs [29]. In particular, superior thermal/chemical/environmental stability,
processability, tailorable permittivity and permeability, corrosion resistance, light
weight, and tunable properties of polyaniline compared to other ICPs remain an
added advantages in EMI shielding applications [30, 31]. However, commercial
utility of ICPs is limited due to its poor mechanical properties. Further, shielding
efficiency achieved in this manner was found to be relatively low due to poor con-
ductivity, leakage of radiation from the microwave-transparent interweave spacing
and absence of any secondary shielding mechanism. As a result, fabrication of other
alternative lightweight shielding materials remains a daunting task. Therefore,
research interests were focused on fabrication of conducting fabrics/textiles as
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 169

protective clothing for human being on exposure to high-frequency electromagnetic


fields. In one such approach, natural or synthetic fibres textiles were coated with
intrinsically conductive polymers via electrochemical and chemical polymerization
during the moulding stage [2]. The other alternative approach involved metalizing
textiles with with Ag, Fe, Co, Ni, Cu, etc. [32]. For example, Ag-plated textiles have
been widely investigated to act as protective clothing exhibiting high shielding effi-
ciency, fastness, excellent electrical conductivity and outstanding thermal stabil-
ity [33].
Therefore, several attempts are continuously being made in developing efficient
ICP-coated fabrics and ICP/metal-incorporated lightweight shielding materials for
the attenuation of EM waves [17–164]. Such textile surfaces covered with ICP/
metal nanoparticle are expected to have a good level of electrical conductivity. In
addition, work has also been made on EMI shielding of ICP incorporated conduct-
ing materials, such as nanocarbon (CNT, graphene), dielectric materials and their
nanocomposites comprising of dielectric/magnetic and dielectric/conducting mate-
rials in EMI shielding. In view of this, the present chapter reviews updated findings
on EMI shielding performance of intrinsically conducting polymers either in pure
form or fabricated by incorporating it on fabrics, metals, such as Ag, Fe, Co, Ni, Cu,
alloy, MWCNT, graphene, iron oxides, BaTiO3, ZnO, TiO2, SiO2 etc.

7.2  Theoretical Aspect of EMI Shielding

Generally, shielding of EM wave is measured by calculating the magnitude of


reduced power or field strength of incident EM wave caused by the shielding mate-
rial. The nature of the shielding material, thickness, distance from the source of EM
wave (r) and frequency of the EM wave are some of most important parameters
accounting for effective shielding. When EM wave are incident on a material, their
barrier towards the propagation could be attributed to the contribution through dif-
ferent mechanisms namely reflection, absorption and multiple internal reflections as
depicted through Fig. 7.2 [17]. Reflection of the EM waves occurs in highly con-
ducting shielding materials (metals, carbon materials) due to presence of mobile
charge carriers, i.e. electrons and holes. Alternatively, EM waves can also undergo
multiple reflections between the opposite surfaces or phases present inside the
shield (internal). The absorption mechanism involved interaction between the elec-
tric/magnetic dipoles of the shielding materials with the electric/magnetic vector of
the electromagnetic radiation incident on the materials. For example, absorption of
the EM waves can take place in materials with high dielectric constant, e.g. BaTiO3,
and high magnetic permeability, e.g. Fe3O4, due to presence of electric dipoles and
magnetic dipoles, respectively.
Generally, materials exhibiting high conductivity, dielectric constant and mag-
netic permeability could account for absorption of EM waves [28]. The presence of
170 S. K. Srivastava

Fig. 7.2 (Modified)
Scheme demonstrating
mechanism of EMI
shielding [17].Reproduced
with permission from
Elsevier

the electric/magnetic dipoles in the shield destroys the electric/magnetic field of the
EM waves by converting it as heat loss. In case of transmission, no attenuation of
EM waves occurs through the shielding material, e.g. glass, polyester and polypro-
pylene. Considering this, understanding of shielding theory is more important and
described briefly as below.
Shielding efficiency (SE) of a material acting as a shield corresponds to ratio of
the magnitude of the incident electric or magnetic field without shielding to electric
or magnetic field with shielding. The total electromagnetic interference shielding
effectiveness (SET) of conducting material is strongly dependent on electrical con-
ductivity, dielectric constant, thickness of materials, shielding efficiency due to
reflection and/or absorption, frequency, affinity to conducting fillers in case of com-
posites, etc. [34].
SET can also be expressed mathematically [24, 25, 32] as

P   ET   HT 
SE T = 10 log10  T  = 20 log10   = 20 log10  
 PI   EI   HI 

where, PI, EI, HI are incident power, electric field and magnetic field intensities,
respectively, whereas, PT, ET, HT are transmitted power, electric field and magnetic
field intensities, respectively.
SET also referred to as electromagnetic interference shielding effectiveness (EMI
SE) of a shielding material can be obtained by summing the individual contributions
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 171

of reflection (R), absorption (A) and multiple reflections (M) in dB [23], i.e. can be
expressed in decibels (dB) [35, 36]:

SE T ( dB) = SE R + SE A + SE M

where SER, SEA and SEM correspond to contributions due to reflection, absorption
and multiple internal reflections, respectively.
SEM can be ignored provided SEA > 10 dB, and correspondingly magnitude of
and total EMI SE become

SE T ( dB) = SE R + SE A

SER and SEA can be written as [23]

SE R ( dB) = 10 log10 (1 − R )
 T 
SE A ( dB) = 10 log10 (1 − Aeff ) = 10 log10  
 (1 − R ) 

The impedance mismatch between the free space and shielding material accounts
for reflection loss (SER). It can be expressed as

 σr 
SE R ( dB) = −10 log10  ,
 16ωε 1ε 0 µ r 

where σΤ, ω, ε1, ε0 and μρ refer to total conductivity, angular frequency, relative per-
mittivity and free space, respectively. This suggests that high SER could be achieved
for a material exhibiting high σΤ, and low ε0 or μT.
Shielding due to absorption (SEA) in dB is related to thickness (t) and skin depth
(δ) of shielding material as [35]
1
1 t  σ ωµ  2
SE A ( dB) = −20 log10 e = −8.68   = −8.68t  T  ,
δ δ   2 

where δ (defined as the distance required by wave to be attenuated to 1/e of its origi-
nal strength)  =  (2/σΤωμ)1/2. This relationship clearly shows that sufficient thick-
nesses, high value of electrical conductivity as well as permeability are the
prerequisite for a good EM wave absorbing material.
The shielding contribution due to multiple reflections (SET) can be expressed [37]:

 σr 
SE T ( dB) = −10 log10  
 16ωε1ε 0 µr 
172 S. K. Srivastava

7.3  E
 lectromagnetic Interference Shielding of ICPs
and ICP-Coated Fabrics

In very early work, conductive polyaniline films have been used in investigating
their performance in electromagnetic interference shielding. However, their poor
mechanical property enables to apply coating of these ICPs by polymerising aniline
and pyrrole on the insulating fabrics for possible EMI shielding response.

7.3.1  ICP-Coated Fabrics

Textile-coated conducting polymers have been employed in applications of electro-


magnetic interference shielding. The choice of conductive polymer coated fabrics
is mainly guided by their flexibility, versatility, low mass and low cost. In view of
this, considerable work has been reported on polyaniline, polyaniline-coated fab-
rics, polypyrrole and pyrrole-coated fabrics for their EMI shielding performance
[38–79].

PANI and PPY-Coated Fabrics

Dhawan et  al. [19] studied shielding behaviour of conducting polyaniline-coated


fabrics in the frequency range 100–1000 MHz and 8–12 GHz. These studies revealed
shielding effectiveness of the order of 30–40 dB and −3 to −11 dB of polyaniline-­
coated fabric corresponding to the frequency range of 100 to 1000  MHz and
8–12 GHz, respectively. UV–Vis–NIR absorption response showed 98% and 94%
energy being absorbed by polyaniline-coated fabric and polyaniline-coated fabrics,
respectively. The effect of controlled doping on electrical properties and permittiv-
ity of para-toluene sulfonic acid (PTSA)-doped polyaniline was investigated [38].
According to this, increasing doping level improved electrical conductivity as well
as complex permittivity improvement followed by electromagnetic radiation block-
ing capacity from −3.8 dB (for emeraldne base) to −23.9 dB corresponding to dop-
ing of 1.0 M PTSA. It was concluded that main phenomenon at low doping level
whereas absorption becomes increasingly important at higher doping levels and
extends dominant contribution towards total attenuation. The electromagnetic inter-
ference shielding efficiency highly electrically conducting polyaniline-camphor
sulfonic acid spin coated on an electrically insulating substrate has been measured
in the frequency range 0.1–1000 MHz [40]. According to this, SE is found to be
>40 dB up to ca. 100 MHz in the near field and 39 dB at 1 GHz in the far field by
using layer structures. The microwave absorbing properties of poly(3,4-­
ethylenedioxythiophene) (PEDOT) microspheres have been investigated in the fre-
quency range of 2–18 GHz [41]. It is noted that maximum reflection loss is achieved
24  dB (thickness: 2  mm) demonstrating excellent performance of hollow
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 173

Fig. 7.3  EMI SE of PANI-coated films obtained from the flanged coaxial test [45]. Reproduced
with permission from Elsevier

microspheres of PEDOT as microwave absorbing material. PANI-coated transpar-


ent thin films as an EMI shielding/absorbing material satisfied EMI SE require-
ments for transparent bodies [43]. The shielding effectiveness of the of hydrochloric
acid-­doped PANI nanofibres composite coating synthesized by an interfacial polym-
erization was found to be in the range of 38–63  dB in the frequency range of
100 kHz–10 GHz. [44].
Kim et  al. [45] obtained free-standing films of emeraldine salt by PANI-­
emeraldine base with dodecylbenzensulfonic acid on a polyethylene terephthalate
film. Figure 7.3 shows variation of EMI SE of the PANI-coated films obtained with
frequency in the range of 30 MHz–1.5 GHz. It was noted that PANI-coated films
(Thickness: 17 ± 2 μm) exhibited overall SE of ~21.6 dB (99.3%) due to the incor-
poration of other intrinsic factors (e.g., permittivity and permeability) together with
sheet resistance. In another work, polyaniline-coated silica fabrics synthesized by
chemical oxidative polymerization of aniline in a protonic acid medium exhibited
resistivity and shielding effectiveness of 20 ± 28 Ohm–cm and 35.61 dB (101 GHz),
respectively [46].
Stempien et al. [48] proposed a very simple reactive ink-jet printing technique
for in situ deposition of polyaniline (and polypyrrole) by chemical oxidation of
aniline hydrochloride (or pyrrole by ammonium peroxydisulfate) on the surface
of textile fabricsaccording to Scheme 7.1. The ink reservoir of print-head 1, ink
reservoir of print-head and oxidant/monomer molar ratio consist of aq. 0.2–1.2 M
aniline hydrochloride or 0.3–0.89 M pyrrole, aq 1:1.2 for PANI (and 1:1 for PPy)
respectively. Figure  7.4 shows variation of EMI SE with frequency
(30 MHz–1.5 GHz) and the number of conductive polymer layers (a/PANI/PAN
composites, b/PPy/PAN composites). The results show the average value of EMI
SE reached ca. 11.1 dB for PANI and ca. 2.8 dB for PPy for the single PANI or
174 S. K. Srivastava

Print-head 1
NH HCl or
2
NH
Print-head 2

Monomer
Layer of conductive polymer
(NH4)2S2O2

Oxidant
Textile x
substrate
y

Scheme 7.1  Scheme of PANI or PPy deposition on textiles by reactive ink-jet printing [48]

PPy layer, respectively. It is also noted that average values of EMI SE for un-
doped polymers are in the range of 6–13 dB for PANI/fabric composites and of
1–10 dB for PPy/fabric composites, respectively. Figure 7.4 also shows highest
SE which corresponds to PANI (25 dB) and PPy (9.3 dB) for 5 layers for conduct-
ing polymer,
Yang et al. [49] studied variation of SE with frequency for pristine PANI and the
BF (bagasse fiber)/PANI with core-shell structure composites in the X-band fre-
quency (Fig.  7.5). It is noted that PANI and BF/PANI composite (Thickness:
~0.4 mm) attained EMI SE in the range of 19.4–22.7 dB corresponding to 27.9 and
30.4 dB, respectively. The higher EMI SE of BF/PANI composites could be attrib-
uted to its larger electrical conductivity compared to pristine PANI. Alternatively,
possibility of superior electromagnetic shielding performance of BF/PANI due to
the multiple reflections induced by the special core-shell structure also cannot be
ruled out. The investigation have also been made on the application of cotton fabrics
coated with PANI [50], PANI-TiO2 hybrid-coated cotton fabric [51], PANI-coated
polyester fabric [52], PANO nanofibre-based functional cotton and nylon fabrics
[53], PANI nanofiber-based ink [54] and metalized fabrics coated with PANI [55] in
EMI shielding applications..

7.3.2  PPy and PPY-Coated Fabrics

PPy is also one of the favourable ICP in the electromagnetic interference shielding
due to its good electrical conductivity and outstanding air stability [56]. However,
relatively lesser amount of work on its application in EMI shielding has been
reported in the literature [57–80]. Sodium p-toluenesulfonate-doped PPy films on
insulating epoxy resin substrates showed EMI SE of was found to be ~30 dB over a
wide frequency range from 30 MHz to 1500 MHz [59]. Fauveaux et al [60] carried
out electromagnetic interference studies on conducting polyaniline composites in
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 175

(a) 30
1 layer 2 layer 3 layer 4 layer 5 layer
Shielding efectivaness SE, dB

25

20

15

10

0
0 250 500 750 1000 1250 1500
Frequency, MHz

(b) 12
1 layer 2 layer 3 layer 4 layer 5 layer

10
Shielding efectivaness SE, dB

0
0 250 500 750 1000 1250 1500
Frequency, MHz

Fig. 7.4  EMI shielding effectiveness change vs. frequency and the number of conductive polymer
layers: a/PANI/PAN composites, b/PPy/PAN composites [48]

the radio frequency range X, Ku, and Ka bands is presented. They noted low perco-
lation low threshold below 1% of mass fraction of polyaniline in the blend with the
shielding effectiveness up to 70 dB depending on the sample. Polypyrrole doped
with anthraquinone-2-sulfonic acid (0.027  mol/L) film of 0.54  mm thickness on
textile substrate exhibited maximum shielding effectiveness of 89.9% (18 GHz) in
176 S. K. Srivastava

Fig. 7.5  EMI SE as a


function of frequency for
PANI and the BF/PANI
composites. [49].
Reproduced with
permission from.Amer
Chem Soc

Fig. 7.6  Electromagnetic shielding effectiveness of (a) PPy-coated raw cotton fabrics and (b)
PPy-coated cotton fabrics treated by NaOH/urea system [63]. Reproduced with permission
from Wiley

the frequency range 1–18 GHz [61]. Polypyrrole-coated polyester nonwoven tex-


tiles displayed EMI SE of 37 dB in the frequency range 100–800 MHz [62].
He et al. [63] prepared polypyrrole-coated cotton fabrics via surface microdis-
solution and in situ polymerization of pyrrole [63]. Figure 7.6 variation of EMI SE
of PPy-coated raw cotton fabrics and PPy-coated cotton fabrics treated by NaOH/
urea system in frequency ranges from 1 to 3000 MHz. It is noted that PPy-coated
cotton fabrics exhibited higher SE (15.4–62.9 dB) compared to raw cotton fabrics
(0.22–44.54  dB). The results could be ascribed to the multiple characteristics of
both EM wave absorption and EM wave reflection. In another study, conducting
bacterial cellulose membranes coated with polyaniline membranes displayed higher
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 177

conductivity and improved absorbing properties in the frequency range of 8–12 GHz


[64]. Gahlout and Choudhary [65] reported EMI SE of − 93 to – 108 dB (Thickness:
3 mm) in highly conducting 5-sulfoisophthalic acid monolithium salt-doped PPy/
MWCNT (5  wt%) composites in X-band (8.2–12.4  GHz). Kim et al [66] poly-
meized pyrrole on nylon 6 woven fabrics and measured EMI SE of 5-40 dB in the
range of 15 MHz-13.5 GHz following reflection as dominat mechansim. It was sug-
gested that conductivity and layer array sequence of the conductive fabric detrmine
EMI SE of the composite fabric Jang et al. [66] polymerized PPy by chemical and
electrochemical methods on nylon 6 woven fabrics. It exhibited high electrical con-
ductivity of about 10 S cm−1 and EMI SE in the range of 30 dB over a wide fre-
quency range up to 13.5 GHz. The electromagnetic shielding effectiveness of cotton
yarn coated with PPy by the vapour phase polymerization technique at the 0.4 mol/L
FeCl3 concentration was found to be about 1 dB in the range of 200–800 MHz [67].
PPy polymerized by chemical method on nylon-6 (PPy film thickness: 35–40 μm)
in the frequency range 100–10,000  MHz attained attenuation values reaching a
maximum of about 3 dB (200–3000 MHz) [68]. The coating of polyaniline formed
on poly(ethylene terephthalate) (PET) fabric formed by in situ polymerization of 1
ml of aniline in presence of 5.44 g FeCl3 and 3 ml surfactant showed shielding
efficiency of 21.28 dB [69]. PPY/carbon fibre composite showed EMI SE of ~23 dB
and followed absorption as the dominant shielding mechanism in the frequency
range of 12.4–18 GHz [70].
In another work, chemically and electrochemically EMI SE of PPY polymerized
in sequence on a polyester (PET) woven fabric was studied at different specific
volume resistivity [71]. Figure 7.7 shows variation of EMI SE, absorbances, and
reflectances of PET fabric/PPy composites with various specific volume resistivities
of the PET fabric/PPy composites. It is noted that EMI SE showed with little change
the frequency range of 50  MHz–1.5  GHz under investigation. The magnitude of
EMI SE was maximally enhanced to 36.6  dB corresponding to 0.20  Ohm  cm of
specific volume resistivity. Such increase in EMI SE is accompanied by increased
conductivity and account for reflection as dominant mechanism.

Fig. 7.7  EMI SE,


absorbances, and
reflectances of PET fabric/
PPy composites with
various specific volume
resistivities [71].
Reproduced with
permission from Elsevier
178 S. K. Srivastava

Polypyrrole-chitosan composite films exhibited EMI SE of ~33.9 dB compared


to PPy film without chitosan (~19.8 dB) in the range of 8–12 GHz [73]. The increase
in EMI SE with the increase in electrical conductivity of the films suggested shield-
ing by reflection. In another work, polypyrrole (PPy) or poly(3,4-­
ethylenedioxythiopene) PEDOT coated on the fabrics through chemical/
electrochemical oxidation of pyrrole or EDOT exhibited EMI SE of ~36 dB over the
wide range of frequency up to 1.5 GHz [74]. 3D-PPy/PEDOT (5 wt%) in a wax
matrix prepared by self-assembly prepared method exhibited effective absorption
bandwidth (deeper than −10 dB) of 6.24 GHz at a thickness of 2.5 mm [75]. Several
other studies are also reported on electromagnetic shielding response of ICP coated
fabrics in X band, W band and radiofrequency range [76], polyester yarns with
polypyrrole, stacked polypyrrole-coated nonwoven textiles [77], polypyrrole/poly-
ester composites [78, 79] and reduced graphene oxide-coated Fe3O4@SiO2@poly-
pyrrol [80].
Table 7.1 records electromagnetic shielding interference performance of con-
ducting polymer-coated fabric.

7.4  Metal-Incorporated ICPs

The performance of ICPs in presence of metals, silver, iron, cobalt, nickel and cop-
per [81–100] including alloys evaluated for their performance in electromagnetic
interference shielding materials is described below [101–111].

7.4.1  Silver-Incorporated ICPs

EMI SE of PPy/fabric and AgPd/fabric complexes were found to be in the range of


8 ~ 80 dB depending on the conductivity and the additional Ag vacuum evaporation
[81]. In another work, Lee and coworkers [82] synthesized polypyrrole and AgPd
compounds coated on woven polyethylene teraphthalate (PET) and nonwoven poly-
ester (PE) fabrics. These findings showed EMI SE of PPy/fabric complexes in the
range of 20 ~ 80 dB depending on the thickness and conductivity. The aqueous
dispersions of poly(acrylic acid) (PAA) in PANI and polystyrene sulfonic acid
(PSS) doped PANI was prepared by in situ polymerization of aniline and used for
coating on cotton fabrics [83]. The four-layered shield of these PANI-PAA-cotton
fabric and PANI-PSS-cotton fabric showed shielding effectiveness of −5.2
and −5.7 dB, respectively. This could be attributed to the enhancement of absorp-
tion loss and multiple reflection contribution due to impedance mismatch at
interfaces.
Luo et  al. [84] prepared polypropylene (PP) fabric coated by polydopamine
(PDA) and silver nanoparticles (PP/PDA/AgNPs) composite modified by one layer
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 179

Table 7.1  records electromagnetic shielding interference performance of conducting polymer-­


coated fabric
Electromagnetic
interference Shielding Dominant
Material Range performance mechanism
PANI-coated fabrics [19] 100–1000 MHz 30–40 dB Absorption
Emeraldine base (EB) −3.8 dB (for EB) Reflection (at low
protonated PTSA-doped to −23.9 dB (1.0 M doping level);
PANIs [38] PTSA) absorption (at
higher doping)
PANI-Camphor Sulfonic 0.1–1000 MHz SE is >40 dB up to ca. —
Acid [40] 100 MHz in the
near-field and 39 dB at
1 GHz in the far-field
PEDOT hollow 2–18 GHz Maximum RL reached Excellent
microsphere [41] approximately −24 dB microwave
(thickness: 2 mm) absorbing
PHCl-doped polyaniline 100 kHz–10 GHz SE: 38–63 dB —
nanofiber coatings (higher
than 35%) [44]
PANI-coated films [45] 30–1500 MHz EMI SE: 21.6 dB Absorption
PANI-coated fabrics [46] 101 GHz SE: 35.61 dB —
Silver-plated polyimide 30 kHz–3 GHz Shielding effectiveness: —
fabric initiated by 54–90 dB
polyaniline [47]
PANI/Bagasse Fiber [49] 8.2–12.4 GHz EMI SE: 28.8 dB Absorption
Cotton fabric-coated PANI 6–14 GHz Ave. EM SE: 3.8 dB Aver. absorption:
[50] 6–14 GHz Ave. EM SE: 6.0 dB 48%
Cotton fabric-coated PPY Aver. absorption:
[50] 50%
PANI-coated polyester 8.2–12.4 GHz Transmission Absorption
fabric [52] loss:53–43%
PANI nanofiber-based 8.2–18 GHz EMI SE: 11–15 dB
functional cotton and nylon
fabrics [53]
PANI nanofiber (50 μm 8.2–18 GHz EMI SE: 13 dB Absorption
thick) [57] EMI SE: 17 dB
PANI nanofiber graphite
[54]
Sodium p-toluenesulfonate-­ 30 EMI SE: about 30 dB —
doped polypyrrole film [59] MHz–1500 MHz
40% of PAni dispersed in 50 MHz–1 GHz, SE: 70 dB —
SAN coated onto a 8.2–18 GHz,
fiberglass panel [60] 33–50 GHz
Anthraquinone2-sulfonic 1–18 GHz Maximum SE: 89.9% at —
acid-doped polypyrrole 18 GHz
applied on textile substrates
[61]
(continued)
180 S. K. Srivastava

Table 7.1 (continued)
Electromagnetic
interference Shielding Dominant
Material Range performance mechanism
PPy-coated polyester 100–800 MHz SE: 37 dB Absorber of EM
nonwoven textiles [62] radiation
PPy-coated cotton fabrics 1–3000 MHz SE: 15.4–62.9 dB —
[63]
Bacterial cellulose 8–12 GHz. SE: ~ −5.0 dB
membranes coated with a
high proportion of
polyaniline [64]
5-sulfoisophthalic acid 8.2–12.4 GHz EMI SE: –93 to −108 dB Absorption
monolithium salt-doped
PPy/MWCNT (5 wt%) [65]
Polypyrrole–nylon 6 50MHz–13.5 GHz EMI SE: 5–40 dB Reflection
composite fabrics [66]
PPy-coated cotton yarns 300–3000 MHz EMSE: ~1 dB —
[67]
PPy-Nylon-6 [68] 100–10,000 MHz EMI SE: 3 dB at 200 —
and 3000 MHz
PPy/carbon fibers [70] 12.4–18 GHz EMI SE: 23 dB Absorption
PET fabric/polypyrrole 50 MHz–1.5 GHz EMI SE; 36 dB (over a Absorption
[71] wide frequency range up
to 1.5 GHz)
PPy-chitosan [73] 8–12 GHz EMI SE: ~33.9 dB Reflection
Polypyrrole/Textile, 50 MHz–1.5 GHz EMI SE: 36 dB up to Reflection
PEDOT on the fabrics [74] 1.5 GHz
5 wt% 3D-PPy/PEDOT/ 2–18 GHz Effective absorption Absorption
wax (thickness: 2.5 mm) bandwidth (deeper than
[75] −10 dB) of 6.24 GHz
(Thickness: 2.5 mm)
Polyaniline-coated fabrics  100–1000 MHz SE: 35.61 dB (101 GHz) —
[76]
Polyester yarns with PPy 200–3000 MHz 2 dB —
[77]

of polydimethylsiloxane (PDMS) (referred to as PP/PDA/AgNPs-C/PDMS-T,


where C and T refer to wt% of silver trifluoroacetate and the time (min) of the com-
posite fabric dipped in the PDMS solution respectively). Figure 7.8a shows EMI SE
of the PP/PDA/AgNPs-C/PDMS-T composite fabrics with different concentrations
in the X band. It is noted that SE increases from 1.9 dB for PP/PDA/AgNPs-5%/
PDMS-40 to 71.2 dB for PP/PDA/AgNPs-25%/PDMS-40. In all probability, porous
structure and high conductivity of the fabric (81.2  S/cm) account for the such
enhancement in the shielding effectiveness of PP/PDA/AgNPs-25%/ PDMS-40.
Figure 7.8b shows variation of SE total and SE absorption and SE reflection of the
PP/PDA/AgNPsC/PDMS-40 at 8.2 GHz. Though SEA and SER both continued to
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 181

(a) (b)
80
70 SE
70 R
25wt% 60 SEA
60 SET
20wt% 50
EMI SE (dB)

EMI SE (dB)
50
40
40 16wt%
30
30 12wt%
20
20
8wt% 10
10
5wt% 0
0
8 9 10 11 12 5 8 12 16 20 25
Frequency (GHz) STA Concentration (wt%)

Fig. 7.8 (a) EMI SE of the PP/PDA/AgNPs-C/PDMS-40 in the X band. (b) SE total, SE absorp-
tion, and SE reflection of the PP/PDA/AgNPs-C/PDMS-40 at 8.2  GHz [84]. Reproduced with
permission from ACS

increase with Ag precursor concentration, SEA exhibits a greater increase compared


to SER Thus, SEA (56.5 dB) of the composite fabric was found to be significantly
higher compared to the SER (12.8 dB) at 8.2 GHz for the Ag precursor concentration
of 25  wt%. These findings are clear indication of reflection acting as dominant
mechanism in the attenuation of electromagnetic waves. The electromagnetic
shielding measurements on silver-plated polyimide fabric initiated by polyaniline
achieved shielding effectiveness in the range of 54–90 dB (30 KHz-3 GHz) [85]. It
also exhibited excellent anti-corrosion resistance, tensile strength, thermal stability,
washing and rubbing fastness.
Fang et al. [88] prepared layer and plane structured composite films comprising
free-standing silver nanowires (AgNWs) and PANI prepared by two-step casting
process and direct-mixing process. It was observed that Ag-NWs form a conductive
network in the layered structure than in the plain structure. As a consequence, elec-
trical conductivity of layer-structured composite prepared via the two-step casting
method is higher compared to the plain-structured composite prepared via the
direct-mixing method. Figure 7.9 shows EMI SE to be above 50 dB over a wide
bandwidth of 0.4 GHz for layer-structured composite films in comparison to plain-­
structured composite film.
Polyaniline microtubes synthesized for the application on silver nanorods
showed reflection loss of about –15.5 dB corresponding to 2400 MHz frequency
[89]. EMI SE of polyaniline filled with graphene, graphene decorated with silver
nanoparticles (Ag@graphene) and graphene decorated with nickel nanoparticles
(Ni@graphene) have been studied in the frequency range from 0.45 to 1.5 GHz and
corresponding finding are displayed in Figure 7.10 [90]. It is noted that enhance-
ments in SET of pristine polyaniline, polyaniline/graphene (0.5  wt.%), Ag@gra-
phene (0.5 wt.%), and Ni@graphene (0.5 wt.%) correspond to 14.5, 15.76, 20.17,
and 17.15  dB, respectively. The conductivity and EMI SE of PAni composite
increased with filler loadings. These findings also showed absorption as the primary
182 S. K. Srivastava

Fig. 7.9  The EMI SE in the X-band frequency range for the Ag-NW/PANI composite films pre-
pared via (a) a two-step casting and (b) a direct-mixing method [88]. Reproduced with permission
from RSC

35
(b)
(a) (b) SET 5.0 wt%
Electrical Conductivity (S/cm)

10 30
(c) SER
25 3.0 wt%
Highest EMI SE (dB)

1.0 wt%
1 (a) 0.5 wt%
20

15
0.01
10
(a) graphene/PAni
(b) Ag@graphene/PAni 5
(c) Ni@graphene/PAni
1E-5 0
(a) (b) (c) (d) (e) (f) (g) (h) (i) (j) (k) (l) (m)
0 0.5 1 3 5
Loading (wt%) Nanocomposites

Fig. 7.10 (A) Electrical conductivity of the PAni and the PAni composites filled with graphene,
Ag@graphene, and Ni@graphene; (B) highest EMI SE of: (a) PAni; PAni composites filled with
graphene, Ag@graphene, and Ni@graphene for 0.5 wt.% (b–d), 1.0 wt.% (e–g), 3.0 wt.% (h–j),
and 5.0 wt.% (k–m) [90]. Reproduced with permission from Elsevier.

shielding mechanism due to the high permittivity of the composites. The highest
EMI SE of 29.33 dB among all samples is achieved by PANI loaded 5.0 wt.% Ag@
graphene.
Polyaniline hollow microspheres (PnHM) nanocomposites of Ag (PnHMAg) by
emulsion polymerization of aniline and Tollen’s reagent as a source for Ag nanopar-
ticles and studied for the variation of transmission loss (T), reflection coefficient (R)
absorption coefficient (A) with frequency corresponding to the different thickness
of the PnHM and PnHMAg in the S and X bands in Figs.  7.11, 7.12, and 7.13
respectively [17]. It is also noted that transmission loss (T) of the PnHM and
PnHMAg decrease with increasing thickness with T being lowest in PnHM corre-
sponding to −21.4 dB (7 GHz) and −23.46 dB (8.2 GHz) in S- and X-bands, respec-
tively. However, no such improvements are noticed in PnHMAg samples in S-bands.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 183

Fig. 7.11  Variation of reflection coefficient of (a) PnHM and (b) PnHMAg for different thickness
at S- and X-band frequencies [17] Reproduced with permission from Elsevier

Fig. 7.12  Variation of absorption coefficient of (a) PnHM and (b) PnHMAg for different thick-
ness at S-and X-band frequencies [17]. Reproduced with permission from Elsevier.

Fig. 7.13  Variation of EMI shielding (SE) of (a) PnHM and (b) PnHMAg for different thickness
at S- and X-band frequencies [17]. Reproduced with permission from Elsevier
184 S. K. Srivastava

The reflection coefficient of PnHM and PnHMAg showed increasing trend with
increasing thickness in S- and X-band regions. In addition, a dip between 8 and
10 GHz appears in PnHMAg samples more likely due to maximum reflection of
electromagnetic radiation by high conducting Ag nanoparticles. The maximum val-
ues of SE in PnHMAg and PnHM correspond to be 19.5  dB (11.2  GHz), 12  dB
(8.5 GHz), respectively. These findings clearly signify enhancement in the shielding
property of polyaniline in presence of conducting Ag nanoparticles.
Motivated by this, Panigrahi and Srivastava [92] prepared hollow polypyrrole
(HPPy)/Ag nanocomposites in the similar manner and explored its electromagnetic
shielding property. Figure 7.14(a) displays the variation of EMI (SE) of PPyHPPy,
HPPy/Ag-2 (2 wt% Ag), HPPy/Ag-5 (5 wt% Ag) and HPPy/Ag-10 (10 wt% Ag) in
the frequency range 0.5–8 GHz. This suggest corresponding values of EMI SE to be
~20–5, ~34.5–6, ~36.5–11.5, ~55.78–20 and ~59–23 dB, respectively. This clearly
indicated relatively higher EMI (SE) values of HPPy and its Ag-loaded nanocom-
posites compared to PPy. This could be attributed to the simultaneous contribution
of internal reflection as well as reflection from outer surface according to the
Fig. 7.14(b).
According to Mu et al. [93], silver-plated PET fabric achieved SE in the range of
50–90 dB in the frequency range of 30 kHz to 3 GHz. The lightweight P ­ Py/polydo-
pamine (PDA)/ AgNWs (0–50 wt%) composites exhibited conductivity in the range
of 0.01–1206.72  S  cm−1 and EMI SE values in X-band varying from and 6.5 to
48.4 dB [94]. EMI SE of emeraldine base form of polyaniline/Ag powder doped
with hydrochloric acid composite (Thickness: ∼70 μm) corresponds to ∼46 dB in
the frequency range of ~107–109 Hz [95, 96].
EMI SE value of 13–15.5 dB is achieved in PPy/Ag/PET nonwoven composite
in the frequency range of 15–3000  MHz with absorption as dominant shielding
mechanism [98]. The flexible thin 20 wt% of silver nanowires embedded into PPy
improved the shielding effectiveness from 6.27  dB (PPY) to 22.38  dB (PPY/
AgNWs) [99]. Such remarkable enhancement in SET was attributed to the substan-
tial enhancement in electrical conductivity in PPY/AgNWs(62.73 S/cm) compared
to PPy (PPY:0.02 S/cm). EMI SE of multilayer structured cuprammonium fabric/
polypyrrole/copper (CF/PPy/Cu) composite ranged between 30.3 and 50.4 dB in the
frequency range of 30-1000 MHz [100].

7.4.2  Fe-, Co- and Cu-Incorporated ICPs

Studies are reported also reported on the application of elemental magnetic metal
incorporated ICPs in the electromagnetic shielding [101–107]. Cobalt-coated poly-
aniline/fly ash cenosphere flexible film of thicknesss 89 ± 3 μm achieved EMI SE of
∼30  dB over the frequency range of 12.4–18  GHz with absorption as dominant
shielding mechanism [102]. Ni/polyaniline-coated polytrimethylene-terephthalate
knitted fabric prepared by in situ chemical polymerization and electroless nickel
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 185

Fig. 7.14 (a) Variation of EMI shielding of PPy, HPPy, HPPy/Ag-2, HPPy/Ag-5, HPPy/Ag-10
with varying frequency at 0.5–8 GHz, (b) Trapping mechanism of EM wave through enhanced
internal reflection in HPPy/Ag: An anticipated scheme [92]. Reproduced with permission from
Nature Springer

plating exhibited significantly high electromagnetic shielding (SE > 40 dB) [103].


In another work, average SE of Ni coated on fly ash cenosphere (FAC) thin film of
thickness 59 ± 4 μm and 133 ± 4 μm films in the frequency range of 8.2–12.4 cor-
respond to 38 and 60 dB SE GHz, respectively [104]. In all likelihood, it is due to
the presence of magneto-dielectric Ni-FAC and increase in conductivity and
186 S. K. Srivastava

permeability of the composite film. Co/PPy (30 wt% in a paraffin matrix) showed


reflection loss <−10 dB located at 11.7–16.47 GHz with a thickness of 2 mm and
with a maximum reflection loss of ~−33 dB (13.6 GHz).

7.4.3  Alloys

The alloying between the alloy filler and the matrix increases the interfacial polar-
ization, saturation magnetization, permittivity, etc., which cannot be obtained in
pure metals. Zhao et  al. prepared fabric-supported PANI/Co-Ni layer structured
composites by the successive coating of polyaniline and Co−Ni magnetic alloy
(loading: 3.99 mg cm−2) with polyaniline (loading: of 2.86 mg cm−2) according to
Scheme 7.2. Figure 7.15 shows frequency dependence of EMI SE of polyaniline-­3-
coated, Co−Ni-30-coated, and polyaniline/Co−Ni-30-coated fabric samples and
quantitative EMI SE enhancement effect within the X-band frequency. These results
demonstrated higher EMI SE by PANI/Co−Ni-30-coated fabrics (SET:
33.95–46.22 dB) compared to either the PANI-coated fabric or the Co−Ni-coated
fabric. This also indicated its capability of >99.9% EM-wave attenuation within the
X-band frequency compared to their single peers (polyaniline-coated fabric and
Co–Ni-coated fabric) or even the sum of them. In another work, Kamchi and
coworkers [110] reported reflection loss values of –22 dB (at 9.52 GHz) and –20.7
dB (at 14.7 GHz) in PANI/PTSA/FeNi/Epoxy corresponding to its thickness of 9.7
and 6.5 mm respectively, dispersed polyaniline-doped camphor sulfonic acid,
carbon-­coated cobalt and FeNi nanoparticles in polyurethane. The composite fabri-
cated in this 8–18 GHz frequency band.

Scheme 7.2  Schematic of the fabrication of layered fabric-supported PANI/Co–Ni compositesa.


Scheme a:Enlarged view of the interlaced fabric structures was shown as the round optical micro-
graphs taken utilizing an optical microscope (BELONA, XSP-OO). Reproduced with permission
from ACS
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 187

Fig. 7.15  Frequency dependence of EMI SE of PANI-3-coated, Co−Ni-30-coated, and PANI/


Co−Ni-30-coated fabrics (above); quantitative EMI SE enhancement effect within the X-band
frequency (below). Reproduced with permission from ACS

Table 7.2 records electromagnetic shielding interference performances of metal


incorporated in fabrics.

7.5  Nanocarbon-Based ICP Nanocomposites

In recent years, nanocarbon-based materials have been used as conductive fillers in


fabrication of ICP nanocomposites for their applications in EMI shielding [111–
113]. Their performance strongly depends on the type of carbonfiller, conductivity
threshold, etc. In view of this, several studies have been made on evaluation PANI/
MWCNT, PPY/MWCNT, PANI/Graphene and PPY/Graphene [101, 114–139].

7.5.1  PANI/CNT Composites in EMI Shielding Applications

Mostaani et al. [114] observed maximum SE value (~60 dB) in PANI/MWCNT


nanocomposites (7 wt%) corresponding to the frequency range of 8.2–12.4 GHz
with absorption as dominant mechanism. In another work, β-naphthalene sulfonic
acid doped polyaniline PANI composites consisting MWCNTs, carbon fibre and
reduced graphene oxide were prepared by chemical oxidative polymerization
route and determined its EMI SE in the frequency range of 8.2–12.4 GHz [115].
The maximum SE was shown by PANI/RGO (39 dB) compared to PANI/MWCNT
(37 dB) and PANI/carbon fibre (31 dB). It was suggested that presence of oxygenic
functional groups in RGO might increase the EM wave attenuation performance of
188 S. K. Srivastava

Table 7.2  Electromagnetic shielding interference performances of ICP/metal and ICP/metal on


a fabric
Electromagnetic
interference Shielding Dominant
Material Range of study performance mechanism
Ag/PPy/-AQSA/AgPd/PE/Ag 1 ∼80 dB
[81] MHz–1.5 GHz
PPy-Anthraquinone-2-sulfonic 1 MHz–1 GHz ~22 dB —
acid (AQSA)-woven PET fabric ~55 dB
[82] ~80 dB
Ag-Nonwoven polyester fiber
(PE)-Ag [82]
Ag-PPy-AQSA-Ag-Pd-PE-Ag
[82]
PANI-PAA-cotton fabric [83] 12.4–18.0 GHz SE: −5.2 dB —
PANI-PSS-cotton fabric [83] 12.4–18.0 GHz SE: −5.7 dB
PP/PDA/AgNPs-25%/ 8.2−12.4 GHz EMI SE:71.2 dB Reflection
PDMS-40 [84]
Ag-plated polyimide fabric 30 kHz to SE: 54–90 dB —
initiated by polyaniline [85] 3 GHz
Polypyrrole–Nylon 6 composite 50 EMI SE: 5–40 dB Absorption
fabrics [86] MHz–13.5 GHz (<20 dB),
Reflectance (>
20 dB)
Ag nanowire (14 vol%)/PANI 2–12.4 GHz EMI SE above 50 dB over —
[88] a wide bandwidth of
0.4 GHz
PANI composite containing 0.45–1.5 GHz EMI SE: 29.33 dB at 1.5 Absorption
5.0 wt.% Ag@graphene [90] GHz
PANI/flower-like copper 300 kHz–3 GHz EMI SE: −45.2 dB (2.78 Absorption
monosulfide (thickness: 3 mm) GHz);
[91] Improved shielding
efficiency below −18 dB
from 300 kHz to 3 GHz
Hollow PPy microsphere/Ag 0.5–8 GHz EMI SE: 59–23 dB Internal
(10 wt%) [92] reflection
Ag plated on PET fabric [93] 30 kHz–3 GHz SE: 50–90 dB Mechanism
changed from
absorption to
reflection with
Ag loading
PPy/PDA/Ag nanowires 8.0–12.0 GHz
EMI SE: 6.5 (0 wt% —
(AgNW) [94] AgNW) to 48.4 dB at
8.0 GHz. (50 wt %Ag
NW)
Emeraldine base form of PANI/ 10 MHz–1 GHz SE: about 46 dB —
Ag powder doped with
hydrochloric acid (thickness:
70 μm) [96]
(continued)
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 189

Table 7.2 (continued)
Electromagnetic
interference Shielding Dominant
Material Range of study performance mechanism
CNT/metal nanoparticle 200–1000 MHz SE: 15–40 dB —
shielding (Thickness:
100–200 μm.fabrics) [97]
PPy/silver/poly(ethylene 15–3000 MHz EMI SE: 13–15.5 dB Absorption
terephthalate) [98]
Cuprammonium fabric/ 30–1000 MHz EMI SE: 30.3–50.4 dB Absorption/
polypyrrole/copper reflection
composite [100]
Co-coated PANIfly ash [102] 12.4–18 GHz EMI SE of ∼ 30 dB Absorption
Ni/PANI-coated PTT knitted 15–3000 MHz EMI SE> 40 dB —
fabric [103]
Fly ash/nickel/PANI [104] 8.2–12.4 GHz SE: 38 dB (59 ± 4 μm), 60 —
dB (133 ± 4 μm)
Co/PPy (30 wt% in paraffin 1–18 GHz RLmax: −33 dB at —
matrix (thickness: 2 mm) [107] 13.6 GHz (Thickness: 2
mm)
Ni/PPy (50 wt%) [108] 8.2–12.4 GHz SET  =  77.87 dB —
[102] PANI/Co–Ni coating on 8.2–12.4 GHz EMI SE: 33.95–46.22 dB —
fabrics [109]
PANI‐PTSA/FeNi/epoxy resin 8–18 GHz RL:−22 dB (9.52 GHz) for —
[110] 9.7 mm,
RL:−20.7 dB (14.7 GHz)
for 6.5 mm

PANI/RGO composite. PANI/graphene/amine functionalized MWCNT hybrid


thin film exhibited highest conductivity of 0.11 S/cm and maximum EMI SE of
21 dB in the frequency range 12–18 GHz [118]. In another report, Au-MWCNT/
PANI nanocomposites prepared by in situ polymerization exhibited total shielding
effectiveness of −16 dB [averaged over the X-band (GHz)] and a minimum reflec-
tion loss of −56.5  dB [119]. Such significantly high value of total SE could be
explained considering its higher dielectric properties and the high electrical
conductivity.
Gupta et al. [120] modified PANI by natural graphite flakes (NGF) to improve
the electrical conductivity followed by incorporating different weight fractions of
MWCNTs to it. This modified PANI was ball milled in presence of MWCNTs for
several hours to generate in situ graphene to form multiphasic composite.
Figure  7.16 shows variation of frequency dependence of shielding effectiveness
due to absorption (SEA), reflection (SER), total (SET) of composites on frequency
and comparison of theoretical and experimental shielding effectiveness (SER and
SEA) for the composites with respect to frequency. It is noted that total EMI-SE of
this composite increases with increasing the MWCNTs content due to the
190 S. K. Srivastava

Fig. 7.16  Dependence of shielding effectiveness due to (a) absorption (SEA), (b) reflection (SER),
(c) total (SET) of composites on frequency and (d) comparison of theoretical and experimental
shielding effectiveness (SER and SEA) for the composites with respect to frequency [120]

synergetic effect of functionalized MWCNTs, graphene and PANI. The multiphase


composite attained maximum value of −98 dB at 10 wt% loading of MWCNTs
with SET dominated by the absorption phenomena compared to the reflection.
Further, theoretical and experimental values of SEA and SER are found to be in
good agreement with one other. The highly conducting PANI coated MWCNTs
prepared by in situ polymerization method exhibited EMI SE of 47.03 dB due to
synergistic combination of the conductive components following absorption as
dominant shielding mechanism [122]. Cao et  al. [123] compared microwave
absorption and effective band width of wax composites comprising Fe3O4/MWCNT
and PANI/Fe3O4/MWCNT. Figure 7.17a shows variation of reflection loss of Fe3O4/
MWCNT and PANI/Fe3O4/MWCNT in wax composites of different thicknesses in
the frequency in the range of 2–18  GHz. It is clearly inevitable that maximum
reflection loss in Fe3O4/MWCNT composites attained∼75 dB (3 mm in thickness).
Figure 7.18b suggests that PANI/Fe3O4/MWCNT/wax to be less effective in micro-
wave absorption compared to Fe3O4/MWCNT/wax composites. Such excellent
microwave absorption performance of Fe3O4/MWCNT/wax could arise from the
dielectric and magnetic loss contributions. Polystyrene microsphere/PANI/
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 191

Fig. 7.17  Reflection loss


plots of (a) Fe3O4/
MWCNT and (b) PANI/
Fe3O4/MWCNT wax
composites [123].
Reproduced with
permission from Amer
Chem Soc

MWCNT [124], PANI/MWCNT/polystyrene [125] PANI/SWCNT [126], PANI/


graphene/MWCNT [120], para-toluene sulfonic acid-doped PANI-graphene nano-
platelets [127] and PANI/graphene [128] have also been investigated for their per-
formance EMI shielding.
PPY-MWCNT-Ag composites produced via UV-reduction showed higher elec-
trical conductivity and SE in comparison to that formed by chemical reduction
[101]. The effect of oxygen plasma treatment of MWCNT has been investigated on
electromagnetic interference shielding of PPY-coated carbon nanotubes [129]. It
was observed that average EMI SE of PANI deposited on MWCNT (plasma treated)
increased to 28.3 dB compared to without oxygen plasma treatment (21.5 dB) and
followed absorption as the main mechanism of EMI shielding. Kim and others
[130] observed remarkable increases EMI SE (~28.6  dB) in PPy coated on
oxyfluorination-­treated MWCNT.
192 S. K. Srivastava

(a) 12
(A) PFF11
21 (A) PFF12
(A) PFF13
(A) PPy
18 10

15
8
(R) PFF11
12

SER (dB)
SEA (dB)

(R) PFF12
(R) PFF13
(R) PPy 6
9

6 4

3
2
12 13 14 15 16 17 18
Frequency (GHz)
(b) 24 3.5
(σ ) PFF11
s
(σs) PFF12
21
microwave conductivity (σs) (S/m)

(σ ) PFF13
s 3.0
18 (σ ) PPy
s

skin depth (δ) (mm)


15 2.5

12
(δ) PFF11
2.0
(δ) PFF12
9 (δ) PFF13
(δ) PPy
6 1.5

3
1.0
0
12 13 14 15 16 17 18
Frequency (GHz)

Fig. 7.18 (a) Dependence of shielding effectiveness (SEA and SER) of PPy and PFF composites as
a function of frequency, showing the effect of ferrofluid concentration on the SE value of the nano-
composites for sample thickness d ∼2.0 mm. (b) Variation of microwave conductivity and skin
depth of PPy and PFF composites as a function of frequency [143]
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 193

7.5.2  P
 ANI/Graphene and PPy/Graphene Nanocomposites
in EMI Shielding Applications

PANI/graphene (free standing) hybrid films prepared by a solution intercalation


showed reflection dominated SET achieving up to 42 dB (SER:32 dB) and 32 dB
(SER: 24.2 dB) in the frequency range of 4–8 GHz and 8–12 GHz, respectively
[131]. The contributions from reflection of 32 and 24.2 dB to this total EMI SE
was found to be very high compared to that of absorption. It was suggested that
high electrical conductivity and larger surface area of the hybrid film could account
for the observed reflection dominant shielding mechanism. Modak et al. [132] also
studied effect of functionalized graphene loading (1, 3 and 5 wt%) in polyaniline
on its EMI SE. It was noted that EMI SE of nanocomposites increased in the fre-
quency range of 2–12 GHz with increasing graphene loading in polyaniline due to
high conductivity of graphene. The maximum value of EMI SE obtained in 5%
(w/w) graphene loaded PANI composite is found to be between 51 and 52 dB in
the above frequency range. Fe3O4/GNPs-NH-PANI) composites prepared by in
situ polymerization/hydrothermal reaction possessed maximum RL value of
absorption of −40.31 dB (thickness: 2.6 mm) [133]. The minimum reflection loss
up to −60.3 dB (11.1 GHz) with sample thickness of 2.86 mm and the effective
absorption bandwidth less than −10  dB (90% microwave absorption) was dis-
played by reduced graphene oxide covalently grafted polyaniline (30 wt%) in par-
affin matrix [134]. It is anticipated that covalent bonds between reduced graphene
oxide nanosheets and PANI nanorods facilitate the formation of the highly con-
ductive networks and account for the absorption of electromagnetic wave. Hu
et  al. [135] measured EMI SE of the samples in the frequency range of 300
KHz–3 GHz for 5, 10, 20 and 30 wt% of the PANI/CuS/RGO (20 wt% CuS/RGO)
composites in paraffin wax (thickness:3 mm). It was observed that EMI SE of the
samples improved to −18.0 dB for 20 wt% of the filler in wax. PPY-coated carbon
fiber@graphene exhibited RLMin, corresponding to −45.12 dB (7.9 GHz) compared
to bare carbon fiber@graphene of −30.53 dB (RLmin.:14.6 GHz) [136]. In another
study, shielding effectiveness of sodium lauryl sulphate-doped polypyrrole
(SLSDPPy), SLSDPPy-MWCNT, SLSDPPy-graphene and SLSDPPy-hybrid car-
bon composites Ku-band correspond to ~ –28.8, –40.6, –53.3 and –79.9 dB respec-
tively [137]. The highest enhancement of SET in SLSDPPy-hybrid carbon
composites could be attributed to the huge surface area offered by microporous
hybrid carbon structures and presence of highly electrically conductive networks.
The sponge-like PPy/RGO aerogel-­based composite at 10  wt% filler loading
achieved effective electromagnetic absorption bandwidth (below −10  dB) of
6.76 GHz, and the highest reflection coefficient of −54.4 dB at 12.76 GHz [138].
Table 7.3 display EMI shielding performance data on nanocarbon/ICP
nanocomposites
194 S. K. Srivastava

Table 7.3  EMI shielding performance data on PANI/Nanocarbon, PPy/Nanocarbon, PANI/


Graphene, and PPy/Graphene
Electromagnetic interference Dominant
Material Range Shielding performance mechanism
PANI/MWCNTs (7 wt%) 8.2–12.4 GHz SE: 60 dB Absorption
[114]
β-naphthalene sulfonic acid 8.2–12.4 GHz SE: PANI/MWCNT: 37 dB), —
(b-NSA)-doped PANI/Carbon SE: PANI/CF : 31 dB)
fillers [115] SE: PANI/rGO : 39 dB)
PANI/MWCNT [117] 12.4–18.0 GHz Shielding effectiveness: Absorption
−27.5 to −39.2 dB
PANI/Graphene/functionalized 12-18 GHz EMI SE: 21 dB Absorption
MWCNT [118]
Au-MWCNT/PANI [119] 8–12 GHz SET: −16 dB, Min. SEA ≫ SER
RL:−56.5 dB
MWCNT 12.4–18 GHz EMI SE: −98 dB Absorption
(10 wt%)–Graphene–
PANI [120]
Polyaniline-coated MWCNT/ 800 SE: 34.1 dB Synergetic
maghemite [121] MHz–2.5 GHz effect of
reflection and
absorption
PANI-coated MWCNT 800 MHz–3 GHz EMI SE: 47.03 dB Adsorption
(oxyfluorinated) [122]
Polystyrene microsphere/ 8.2–12.4 GHz EMI SE; ∼22.7–23.2 dB —
PANI/MWCNT[124]
PANI/SWCNT (25 wt%) [126] 2–18 GHz EMI SE: 31.5dB Absorption
PANI/Graphene (33 wt%) 2–18 GHz EMI SE: 34.2 dB Absorption
[126]
Para-Toluene Sulfonic Acid 8–12 GHz −14.5 dB Absorption
(p-TSA)-doped PANI–
Graphene nanoplatelets
(10 wt%) [127]
PANI/Graphene (33 wt%) 2–18 GHz RLmax = −45.1 dB —
[128] (Thickness: 2.5 mm)
COOH-MWCNT [101] 4.7–7.7 GHz 21 dB —
MWCNT-PPy [101] 23 dB
PPy-MWCNT-Ag (Chem. 27 dB
red.) [101] 30 dB
PPy-MWCNT-Ag (UV-red.)
[101]
PPy-coated carbon nanotubes 800 MHz– Average EMI shielding Absorption
[129] t3 GHz efficiency of PPy-coated
MWCNTs increased from
21.5 to 28.3 dB
(continued)
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 195

Table 7.3 (continued)
Electromagnetic interference Dominant
Material Range Shielding performance mechanism
PPy-coated 800 MHz–3 GHz EMI SE: 28.6 dB Absorption
MWCNT(oxyfluorinated)
[130]
Polyaniline/Graphene [131] 4–12 GHz SET ∼42 dB (4–8 GHz), Reflection
SET ∼32 dB (8–12 GHz)
PANI/Functionalized graphene 2–12 GHZ 51–52 dB Absorption
(5%) [132]
Covalently bonded GNPs-NH-­ 2–18 GHZ Max. RL: −40.31 dB Absorption
PANI nanorod modified by (Thickness: 2.6 mm) and
Fe3O4 [133] corresponding bandwidth
with effective attenuation
(RL <10 dB) is up to
9.62 GHz (7.85–17.47 GHz)
rGO-g-PANI (30 wt%) in 2–18 GHz Min. reflection loss Absorption
paraffin matrix [134] (Thickness: 2.86 mm):
−60.3 dB (11.1 GHz
Effective absorption
bandwidth less than −10 dB
(90% microwave absorption)
is as wide as 4.7 GHz (from
9.3 to 14 GHz)
PANI with 20 wt % of CuS/ 300 KHz–3 GHz EMI SE: up to −18 dB —
reduced graphene oxide [135]
Carbon fiber@Graphene@PPy 2–18 GHZ Max. RL: –45.12 dB at 7.9 Trapping
(20 wt % in wax) GHz (Thickness: 2.5 mm). Absorption
Sodium lauryl sulphate doped 12.4–18.0 GHz −79.9 dB
polypyrrole-Hybrid carbon
composites [137]
S-PPy/RGO aerogel(10 wt%)/ 2–18 GHz Effective electromagnetic Absorption
Paraffin [138] absorption bandwidth
(below 10 dB): 6.76 GHz;
Highest reflection
coefficient: 54.4 dB (12.76
GHz)
MWCNTs/L-PANI [139] 2.0–18.0 GHz RL below −10 dB (5.8–7.4 —
MWCNTs/D-PANI [139] GHz), and max.
value: = −14.95 dB
(6.72 GHz).
RL below −10 dB (5.2–7.0
GHz) and max. value:
−13.11 dB (5.92 GHz)
196 S. K. Srivastava

7.6  I CPs Nanocomposites Consisting Dielectric/Magnetic/


Conducting Materials

An efficient shielding material should possess either mobile charge carriers (elec-
trons or holes) or electric and/or magnetic dipoles to interact with the electric (E)
and magnetic (H) vectors of the incident EM radiation [4]. Therefore, highly con-
ducing and dielectric materials act as an efficient shield in blocking incoming EM
waves. As a result, several investigations have been made dealing with variety of
materials exhibiting on wide range of electrical conductivity (σ), good electromag-
netic attributes (permittivity, ε, and permeability, μ) and their geometries (such as
core/shell). In view of this, application of ICP composites comprising magnetic/
dielectric/conducting materials is studied for their applications EMI shielding
[140–161].
PANI/BaTiO3 composites have shown excellent microwave shielding response
with absorption-dominated total shielding effectiveness (SET) value of −71.5 dB in
12.4–18 GHz [140]. The electromagnetic and microwave absorbing properties of
magnetite nanoparticles decorated carbon nanotubes/polyaniline multiphase (CNTs/
PANI/Fe3O4) heterostructures have been studied in the frequency range of 2–18 GHz
[142]. These results suggested reflection loss of −22.4 dB at 11.7 GHz with about
6.9 GHz bandwidth below −10 dB when used 1.5 g NH2Fe(SO4)2·6H2O in the prep-
aration of this composite. Such microwave absorbing performance has been
accounted with respect to both dielectric and magnetic losses of the composite het-
erostructure. The ferrofluid-based nanoarchitectured polypyrrole composites (PFF)
with different monomer/ferrofluid weight ratios were synthesized according to
Scheme 7.3 [143]. Figure 7.18 shows the variation of the SEA and SER with fre-
quency for PPy and PFF composites in 12.4–18.0 GHz. These results demonstrated
increase in the absorption of microwave and overall shielding effectiveness with the
concentration of the ferrofluid up to 23.5 dB observed in PFF (Py:FF::1:3)
­composites. In all probability, higher dielectric and magnetic losses induced by
Fe3O4 nanoparticles could be attributed to the increase in the absorption with the
addition of ferrofluid in PFF samples.
Singh et  al. [144] studied microwave absorption properties of 3D chemically
modified graphene/Fe3O4(GF) incorporated polyaniline composites in the frequency
range of 12.4–18 GHz. These findings showed higher SEA of 22–26 dB with a SER
of 4.7–6.3  dB for the composites consisting 1:1 and 1: 2 weight ratio of aniline
monomer to chemically modified graphene/Fe3O4. This mainly arises due to the
synergistic effect of graphene and Fe3O4 with polyaniline. In addition, higher dielec-
tric and magnetic losses in the composite could also be another factor responsible
for its higher SE. The porous structure of PVDF/10 wt% loaded Fe3O4 decorated
polyaniline/single wall carbon nanohorn (1  wt%) exhibited high conductivity,
dielectric loss and permeability. [145]. Further, 2 mm thick of this sample achieved
high EMI SE value (−29.7 dB) in 14.5–20.0 GHz frequency range. In another work,
highly flexible porous PVDF/Fe3O4 decorated polyaniline/RGO(5 wt%) composite
conductivity and EMI SE value of ~1.10 × 10−1 S cm−1 and −28.18 dB, respectively
[146]. Poly-(aniline)-coated fabric with incorporated dielectric (BaTiO3) and mag-
netic (Fe3O4) nanoparticles resulted in total shielding (SET) of −16.8 and −19.4 dB,
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 197

Scheme 7.3  Schematic representation for the synthesis of polypyrrole–ferrofluid nanocompos-


ites and the interaction of the microwave with the polymer composite results in attenuation due to
the scattering with the nanoparticle [143]

respectively compared to PANI coated fabric (SET:–15.3 dB) [147]. This is due to
the better matching of input impedance, reduction of skin depth as well as addi-
tional dielectric/magnetic losses. Moon et al. [149] observed synergetic improve-
ment in electromagnetic interference shielding characteristics of polyaniline-coated
graphite oxide/γ-Fe2O3/BaTiO3 nanocomposites.
Polyaniline-TiO2-γ-Fe2O3 nanocomposite synthesized by microemulsion method
showed higher shielding effectiveness due to absorption (SEA  ∼  45  dB) than the
polyaniline-γ-Fe2O3 (SEA  ∼8.8  dB) and polyaniline-TiO2 (SEA  ∼  22.4  dB) [150].
Graphene@Fe3O4@PANI@TiO2 nanosheets containing 50 wt% paraffin (thickness:
1.6  mm) [151] achieved maximum reflection loss up to −41.8  dB (14.4  GHz).
Belaabed et al [154] dispersed polyaniline and Fe3O4 fillers dispersed in epoxy resin
for electromagnetic applications. These study evealed a minimum reflection coeffi-
cient of −42 dB was observed at 16.3 GHz with a thickness around 1 mm for compos-
ites containing 15% of PANI and 10% of Fe3O4 (εʹ = 10). PPy containing multi-layered
graphene anchored on TiO2 (5%) composite have shown total shielding effectiveness
of 53  dB in the frequency range of 12.4–18  GHz [155]. Fe3O4/ZnO, PTSA [152],
polypyrrole-reduced graphene oxide-Co3O4 [153], epoxy-PPy/Fe3O4–ZnO (2:1) [156]
and RGO/PPY/nanotubes/Fe3O4 aerogel [157] have also been investigated in electro-
magnetic shielding interference. Table 7.4 describe EMI shielding performance data
for ICP/Dielectric, ICP/Dielectric/Magnetic and ICP/Dielectric.Conducting materials.
Table 7.4  EMI shielding performance data on ICP/Dielectric, ICP/Dielectric/Magnetic, and ICPs/Dielectric/Conducting Materials
198

Electromagnetic interference Shielding


Material Range performance Dominant mechanism
BaTiO3 [140] 12.4–18 GHz SET: −71.5 dB Absorption
PANI/Silica [141] 8–12 GHz PANI/Silica (75/25): −14.37 dB Absorption
CNT/PANI/Fe3O4 [142] 2–18 GHz RL: –22.4 dB at 11.7 GHz with 6.9 G Hz —
bandwidth below –10 dB
Ferrofluid/PPy/ 12.4–18 GHz SET: 23.5 dB Absorption
Fe3O4 (8–12 nm) [143]
Chemically modified graphene/ 12.4–18 GHz SEA of 22–26 dB Absorption
Fe3O4/PANI [144] SER of 4.7–6.3 dB
PVDF/Fe3O4/PANI/ single wall 100 MHz– EMI SE ≈ −29.7 dB —
carbon nanohorn [145] 20 GHz
PVDF/Fe3O4/RGO/PANI [146] 8.2–12.4 GHz EMI SE value (≈–28.18 dB) Absorption
PANI-coated graphite oxide/γ-Fe2O3/ 800 MHz–3 GHz EMI SE: ~31.5–37 dB Reflection/absorption (synergetic effect)
BaTiO3 [149]
Polyaniline–TiO2–γ-Fe2O3 [150] 12.4–18 GHz SEA ∼ 45 dB Absorption
Graphene@Fe3O4@PANI@TiO2 2–18 GHz RLmax : −41.8 dB, absorption bandwidth of Absorption
[151] RL < −10 dB ~up to 3.5 GHz (Thickness: 1.6
mm)
Fe3O4/ZnO, PTSA as dopant of 8.4–11.6 GHz RL less than −10 dB —
polyaniline [152]
PPy-RGO-Co3O4 in paraffin 50% 2–18 GHz RLmax: −33.5 dB at 15.8 GHz (Thickness: 2.5 mm) —
[153]
S. K. Srivastava
Electromagnetic interference Shielding
Material Range performance Dominant mechanism
PANI (15 %)/Fe3O4 (10 %)/epoxy, 12.4–18 GHz Min. Ref. Coeff.: −42 dB at 16.3 GHz Strong absorption
[154] (thickness ~1 mm)
PANI (15 %)/Fe3O4 (25 %)/epoxy, Min. Ref. Coeff.: −37.4 dB at 14.85 GHz
[154] (thickness ~1 mm)
PPy/Graphene/TiO2 [155] 12.4–18 GHz SET: 53 dB Absorption
15 % Epoxy-PPy/Fe3O4-ZnO (Fe3O4/ 8.2–12.4 GHz RLmax : to –32.53 dB at 9.96 GHz Absorption of microwaves
ZnO = 2:1) [156] (Thickness: 2 mm)
RGO/PPY/nanotubes/Fe3O4 for 2–18 Ghz RLmax : −49.2 dB (11.8 GHz) with an effective Absorption behavior due to the multiple
(thickness: 3.0 mm) [157] absorption bandwidth below −10 dB reaching reflections/polarizations/relaxation
6.1 GHz (9.8–15.9 GHz) processes (absorption > 90%)
Iron oxide/polypyrrole core/shell 1–18 GHz RLmax (absorption): −37 dB at 10.3 GHz
[159]
Core–shell Fe3O4@PPy [160] 2–18 GHz RLmax: −41.9 dB at 13.3 GHz (matching (Absorption)
thickness: 2 mm)
Fe3O4@C@PANI  2–8 GHz EMI SE: 65 dB Absorption
(Fe3O4@C:aniline = 1:9 wt/wt) [161]
Fe3O4@SiO2@PPY (pyrrole/Fe3O4@ 2−8.5 GHz EMI SE: ~32 dB Reflection
SiO2 = 9:1)[162]
5 wt% PANI@nano-Fe3O4@Carboon 8.2–18 GHz EMI SE: 29 dB Enhanced EM wave absorption
fiber in epoxy [163]
N-doped graphene@PANI nanorod 2–18 GHz RLmax: −40.8 dB (14.8 GHz), Absorption —
@Fe3O4 [164] bandwidth with a RL below −10 dB (10.4–
15.5 GHz) with a thickness of 2.7 mm
Heterostructure (10 wt %) 12–18 GHz EMI SE: −40 dB (18 GHz) Absorption
MWCNT(3 wt%) in epoxy [165]
(continued)
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic…
199
Table 7.4 (continued)
200

Electromagnetic interference Shielding


Material Range performance Dominant mechanism
Functionalized l@/Polypyrrole core/ 2–18 GHz −15.2 dB (13.0 GHz) Absorption
shell composites [166]
Ni (core)/polyaniline (shell) with 2–18 GHz RL min: −35 dB at 17.2 GHz (thickness: 5 mm) –
21.6 wt% [167]
γ-Fe2O3/microporous SiO2/ 2–18 GHz F RLmax: −51.24 dB (7.44 GHz) with thickness of Absorption
polypyrrole microspheres 4.0 mm, Effective absorption bandwidth
(14.2 wt%) in wax [168] (RL < −10 dB) 4.16 GHz
Graphene@PANI@porous TiO2 RLmax: Up to –45.4 dB at thickness of 1.5 mm, Multiple reflections
[169] Absorption bandwidths exceeding –10 dB were
11.5 GHz (thickness: 1–3.5 mm)
S. K. Srivastava
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 201

7.7  C
 ore@Shell Materials with Single and Dual Interface
in EMI Shielding Applications

Recently, EMI materials based on core-shell morphology are receiving considerable


attention due to their multifaceted applications [158]. In view of this, interest has
also been focused on ICP-based functional core-shell composites for
electromagnetic interference shielding/microwave absorption. Azadmanjiri et  al.
[159] measured reflection loss of 70 wt% of PPy/silane/magnetic nanoparticles
(MNPs) filled paraffin wax composites prior to heat treatment and after microwave
treatment in the frequency range of 1–18 GHz and corresponding findings are dis-
played in Figure 7.19a and 7.19b respectively. It is noted that PPy/silane/MNPs
composite (thickness: 3.5 mm) prior to microwave treatment showed maximum
reflection loss of –18.0 dB at 7.8 GHz. Interestingly, microwave treated PPy/silane/
MNPs showed maximum reflection loss of –37 dB at 10.3 GHz. Such enhanced
performance of the plasma treated composite compared to untreated one is ascribed
to its higher electrical conductivty of more ordered graphitic-like PPy shell structure.
Core-shell Fe3O4@PPy displayed better reflection loss performance with a wider
effective absorption bandwidth (<−10.0  dB) over the entire Ku band [160]. The
excellent electromagnetic wave absorption properties could be attributed to the
enhanced dielectric loss from the PPy shell in Fe3O4@PPy core-shell composites.
Manna and Srivastava [161] fabricated trilaminar core-shell composites of
Fe3O4@C@polyaniline as schematically shown in Fig. 7.20. The prepared compos-
ites were investigated for its performance in EMI shielding by plotting frequency vs
SEA, frequency vs SER, and frequency vs EMI SE of PFC composites in the fre-
quency range of 2–8 GHz as shown in Fig. 7.21. It is noted that SET follows the
trend as PFC-10 (Fe3O4@C = 9:1): ∼65  dB  >  PFC-20 (Fe3O4@C = 8:2):
∼55 dB > PFC-30 (Fe3O4@C = 7:3): ∼52 dB > FC (Fe3O4@C: ∼41 − 20 dB > Fe3O4 :
∼15 dB. Such high value of shielding efficiency could be ascribed to the presence
of dual interfaces and dielectric-magnetic integration in Fe3O4@C@Polyaniline. In
other words, higher dielectric loss through interface polarization and relaxation
effects in Fe3O4@C@Polyaniline could also contribute toward its superior micro-
wave absorption ability. Alternatively, possibility of more and more electromag-
netic energy converted to microcurrent, trilaminar Fe3O4@C@Polyaniline
core-shell structure due to the increase in impedance matching also cannot be over-
ruled. They also extended their work on Fe3O4@SiO2@PPy and observed highest
total shielding efficiency (∼32 dB) for Fe3O4@SiO2/Pyrrole wt/wt = 1:9 and fol-
lowed reflection as the dominant shielding mechanism [162]. Such performance
was attributed to poor impedance matching between the PPy (conducting)/SiO2
(insulating), high electrical conductivity of Fe3O4@SiO2@PPy and presence of
interface could account such enhancing the total shielding efficiency. Thus EM
interference shielding in Fe3O4@SiO2@PPy and Fe3O4@C@PANI trilaminar
core@shell nanocomposites is controlled by tuning of the shells through switching
of the mechanism as shown in Scheme 7.4.
202 S. K. Srivastava

Fig. 7.19 (a) Reflection loss of PPy/silane/MNPs prior to microwave heat treatment with a con-
centration of 70 wt% in paraffin wax [159]. Reproduced with permission from Springer. (b)
Reflection loss of microwave-treated PPy/silane/MNPs with a concentration of 70 wt% in paraffin
wax [159]. Reproduced with permission from Springer

Movassagh-Alanagh and others [163] preparedepoxy-based hybrid composites


(thickness:1.5  mm) filled with 1  wt% of polyaniline@nano-Fe3O4@CFs (carbon
fiber). It showed a maximum reflection loss (RL) value corresponding to −11.11 dB
with an effective absorption bandwidth of about 6  GHz (frequency range:
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 203

Fig. 7.20  Schematic Presentation of fabrication of Fe3O4@C@PANI ternary composite [161].


Reproduced with permission from ACS

Fig. 7.21 (a) Plots of frequency vs SEA, (b) frequency vs SER, and (c) frequency vs EMI SE of
PFC composites [161]. Reproduced with permission from ACS

8.2–18  GHz). The reflection loss and maximum absorption of the N-doped gra-
phene@ PANI nanorods@Fe3O4 hierarchical structures correspond to −10  dB
(10.4–15.5 GHz) and −40.8 dB (14.8 GHz), respectively [164]. Wang et al. [165]
observed superior microwave absorption N-doped graphen@polyaniline @
Fe3O4nanocluster compared to graphene@Fe3O4in the range of 2 and 18 GHz in all
probability due to impedance matching and interfacial polarization.
204 S. K. Srivastava

Scheme 7.4 Tuning of Shells in Trilaminar  Core@Shell  Nanocomposites in Controlling


Electromagnetic Interference through Switching of the Shielding Mechanism [162]. Reproduced
with permission from ACS

Xu et al. [166] reported that minimum reflection losses of Ni/PPy composites of


Ni/Py ratio of 4:1 and 2:1correspond to −15.2  dB (13.0  GHz) and  −14.8  dB
(14.4  GHz), respectively. They also noted electromagnetic absorption less than
−10 dB in the 11–15.4 GHz and 12–17.5 GHz respectively due to the synergetic
consequence of the Ni cores and PPy shells in Ni/PPY composite. Dong et al. [167]
reported enhanced microwave absorption of Ni/PANI core-shell nanocomposites in
the 2–18 GHzby dual dielectric relaxation and observed absorption properties less
than −10 dB in the frequency range of 4.2–18GHz. Such performance is mainly
ascribed to the EM matching (ferromagnetic Ni core/dielectric PANi shells) and the
enhanced core/shell interfacial relaxation.
Core/shell/shell γ-Fe2O3/microporous SiO2/polypyrrole microspheres (14.2 wt%)
loaded in paraffin wax matrix (Thickness: 4.0  mm) showed maximum reflection
loss of −51.24 dB (7.44 GHz) and effective absorption bandwidth (RL < − 10 dB) of
4.16 GHz [168]. This could be attributed to the synergetic effects, dielectric losses,
magnetic losses and also the microporous SiO2 network. Scheme 7.5 displays the
possible microwave absorption mechanism involved in attenuation of electromag-
netic waves. The response of core-shell graphene@PANI@porous TiO2 composites
have also been studied for shielding electromagnetic applications [169]. Table 7.4
also provide EMI shielding performance data on binary and ternary core@shell
materials.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 205

ve
(a) wa
tted
mi
In
ci ns
de tra multiple reflection
nt
w
av
e SiO2

e absorber wave
av
te dw
flec
Re

Conductivity loss
(b) Interfacial polarization (c)
Interfaces

Dipole
polarization Magnetic loss

SiO2 hopping electrons


γ-Fe2O3/m-SiO2/PPy
migrating electrons

Scheme 7.5 Possible microwave absorption mechanisms of γ-Fe2O3/m-SiO2/PPy MSs. (a)


Multiple reflections and scatterings, (b) interfacial polarization and dipole polarization, and (c) con-
ductivity loss and magnetic loss [168]

7.8  Summary

A suitable shielding material needs to be considering its mechanical flexibility and


resistance towards corrosion. Considering this, shielding materials based on metals
and alloys suffer from certain drawbacks, such as from heavy weight, corrosion
resistance, ease of processing, etc. In view of this, intrinsically conducting poly-
mers, specially polyaniline and polypyrrole are found to be most promising candi-
dates for EMI shielding compared to metals due to their lightweight, noncorrosive
nature andcommercial viability. In view of this, performance of ICPs, especially
PANI and PPy, coated on fabrics and nanocomposites comprising conducting/mag-
netic/dielectric materials have been reviewed for their performance in EMI shielding.

References

1. Chung, D.  D. (2000). Materials for electromagnetic interference shielding. Journal of


Materials Engineering and Performance, 9, 350–354.
2. Chung, D. D. L. (2001). Electromagnetic interference shielding effectiveness of carbon mate-
rials. Carbon, 39, 279–285.
3. Geetha, S., Kumar, K.  K. S., Rao, C.  R. K., Vijayan, M., & Trivedi, D.  C. (2009). EMI
shielding: Methods and materials—A review. Journal of Applied Polymer Science, 112,
2073–2086.
206 S. K. Srivastava

4. Saini, P., & Arora, M. (2012). Microwave absorption and EMI shielding behavior of nano-
composites based on intrinsically conducting polymers, graphene and carbon nanotubes. In
A. D. Gomes (Ed.), New polymers for special applications. InTech: Rijeka, Croatia.
5. Parveen, S., Rana, S., & Fangueiro, R. (2013). A review on nanomaterial dispersion, micro-
structure and mechanical properties of carbon nanotube and nanofiber reinforced cementi-
tious composites. Journal of Nanomaterials. https://doi.org/10.1155/2013/710175.
6. Jagatheesan, K., Ramasamy, A., Das, A., & Basu, A. (2014). Electromagnetic shielding
behaviour of conductive filler composites and conductive fabrics: A review. Journal of Fibre
and Textile Research, 39, 329–342.
7. Joshi, A., & Datar, S. (2015). Carbon nanostructure composite for electromagnetic. Journal
of Pramana – Physics, 84, 1099–1116.
8. Saini, P. (2015). Fundamentals of conjugated polymer blends, copolymers and composites:
Synthesis, properties, and applications. Wiley-VCH.
9. Srivastava, S. K., & Mittal, V. (2017). Advanced nanostructured materials in electromagnetic
shielding. In S.  K. Srivastava & V.  Mittal (Eds.), Hybrid nanomaterials: Developments in
energy, environments and polymer nanocomposites (pp. 241–320). Scrivener-Wiley.
10. Singh, A.  K., Shishkin, A., Koppel, K., & Gupta, N. (2018). A review of porous light-
weight composite materials for electromagnetic interference shielding. Composites Part B:
Engineering, 149, 188–197.
11. Sankaran, S., Kalim, D., Ahamed, M.  B., & Pasha, S.  K. K. (2018). Recent advances in
electromagnetic interference shielding properties of metal and carbon filler reinforced flex-
ible polymer composites: A review. Composites Part A: Applied Science and Manufacturing,
218, 49–71.
12. Maity, S., & Chatterjee, A. (2018). Conductive polymer-based electro-conductive tex-
tile composites for electromagnetic interference shielding: A review. Journal of Industrial
Textiles, 47, 2228–2252.
13. Chhetri, S., Adak, N. C., Samanta, P., Murmu, N. C., Srivastava, S. K., & Kuila, T. (2019).
Synergistic effect of Fe3O4 anchored N-doped rGO hybrid on mechanical, thermal and elec-
tromagnetic shielding properties of epoxy composites. Composites Part B: Engineering, 166,
371–381.
14. Sharma, V., Manna, K., Srivastava, S. K., & Chandra, A. (2019). Hollow nanostructures of
metal oxides as efficient absorbers for electromagnetic interference shielding. Journal of
Physics D, 52, 015301. https://doi.org/10.1088/1361-­6463/aae4f5.
15. Manna, K., & Srivastava, S. K. (2018). Contrasting role of defect-induced carbon nanotubes in
electromagnetic interference shielding. Journal of Physical Chemistry C, 122, 19913–19920.
16. Murugaiyan, P., Mitra, A., Panda, A. K., Santhosh, K. A., Roy, R. K., Manna, K., & Srivastava,
S. K. (2019). Electromagnetic interference shielding effectiveness of amorphous and nano-
composite soft magnetic ribbons. Physica B Condensed Matter, 568, 13–17.
17. Panigrahi, P., & Srivastava, S. K. (2015). Tollen’s reagent assisted synthesis of hollow poly-
aniline microsphere/Ag nanocomposite and its applications in sugar sensing and electromag-
netic shielding. Materials Research Bulletin, 64, 33–41.
18. Lin, J.-H., Li, T. A., Lin, T. R., Jhang, J.-C., Lou, C.-W., Lin, J.-H., Li, T. A., Lin, T. R.,
Jhang, J.-C., & Lou, C.-W. (2019). Processing techniques and properties of metal/polyester
composite plain material: Electromagnetic shielding effectiveness and far-infrared emissiv-
ity. Journal of Industrial Textiles, 49, 365–382.
19. Dhawan, S. K., Singh, N., & Venkatachalam, S. (2002). Shieldingbehaviour of conducting
polymer-coated fabrics in X-band, W-band and radio frequency range. Synthetic Metals, 129,
261–267.
20. Huang, C. Y., Mo, W. W., & Roan, M. L. (2004). Studies on the influence of double-layer
electroless metal deposition on the electromagnetic interference shielding effectiveness of
carbon fiber/ABS composites. Surface and Coatings Technology, 184, 163–169.
21. Yuping, D., Shunhua, L., & Hongtao, G. (2005). Investigation of electrical conductivity and
electromagnetic shielding effectiveness of polyaniline composite. Science and Technology of
Advanced Materials, 6, 513. https://doi.org/10.1016/j.stam.2005.01.002.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 207

22. Vulpe, S., Nastase, F., Nastase, C., & Stamatin, I. (2006). PAN–PAni nanocomposites
obtained in thermocentrifugal fields. Thin Solid Films, 495, 113–117.
23. Cui, G., Lu, Y., Zhou, W., Lv, X., Hu, J., Zhang, G., & Gu, G. (2019). Excellent microwave
absorption properties derived from the synthesis of hollow Fe3O4@reduced graphite oxide
(RGO) nanocomposites. Nanomaterials, 9, 141/1–141/12.
24. Joo, J., & Epstein, A. J. (1994). Electromagnetic radiation shielding by intrinsically conduct-
ing polymers. Applied Physics Letters, 65, 2278–2280.
25. Yangyong, W., & Xinli, J. (2005). Intrinsically conducting polymers for electromagnetic
interference shielding. Polymers for Advanced Technologies, 16, 344–351.
26. Jiang, D., Murugadoss, V., Wang, Y., Lin, J., Ding, T., & Wang, Z. (2018). Electromagnetic
interference shielding polymers and nanocomposites: A review. Polymer Reviews, 59(2),
280–337.
27. Srivastava, S. K. (2019). Rubber/conducting polymer blends: A review. In V. Mittal (Ed.),
Conducting polymer composites (pp. 157–194). Australia: Central West Publishing.
28. Pannigrahi, R. (2015). Fabrication of core@shell structure of conducting polymer micro-
spheres and their applications in environmental remediation. Kharagpur: Indian Institute of
Technology. (Ph.D. Thesis).
29. Yoshida, K. (2005). Part 1: Regular papers, short notes and review papers. Japanese Journal
of Applied Physics, 44, 2025–2029.
30. Pannigrahi, R., Srivastava, S.  K., & Peonteck, J. (2018). Fabrication of elastomer blends.
Involving core (Polystyrene)@Shell (Polyaniline) approach, their characterization and appli-
cations in electromagnetic shielding. Rubber Chemistry and Technology, 91, 97–119.
31. Bhadra, S., Khastgir, D., Singha, N. K., & Lee, J. H. (2009). Progress in preparation, process-
ing and applications of polyaniline. Progress in Polymer Science, 34, 783–810.
32. Topp, K., Haase, H., Degen, C., Illing, G., & Mahltig, B. (2014). Coatings with metallic
effect pigments for antimicrobial and conductive coating of textiles with electromagnetic
shielding properties. Journal of Coating Technology and Research, 11, 943–957.
33. Yu, D., Wang, Y., Hao, T., Wang, W., & Liu, B. (2018). Preparation of silver-plated poly-
imide fabric initiated by polyaniline with electromagnetic shielding properties. Journal of
Industrial Textiles, 47, 1392–1406.
34. Panigrahi, R., Srivastava, S. K., & Pionteck, J. (2018). Fabrication of elastomer blends involv-
ing core (polystyrene)@shell (polyaniline) approach, their characterization and applications
in electromagnetic shielding. Rubber Chemistry and Technology, 91, 97–119.
35. Colaneri, N. F., & Shacklette, I. W. (1992). EM1 shielding measurements of conductive poly-
mer blends. IEEE Transactions on Instrumentation and Measurement, 41, 291–297.
36. Zhang, C.-S., Ni, Q.  Q., Fu, S.-Y., & Kurashiki, K. (2007). Electromagnetic interference
shielding effect of nanocomposites with carbon nanotube and shape memory polymer.
Composites Science and Technology, 67, 2973–2980.
37. Ott, H. W. (2009). Electromagnetic compatibility engineering. New Jersey: Wiley.
38. Saini, P., Arora, M., Arya, S.  K., & Tawale, J.  S. (2014). Effect of controlled doping on
electrical properties and permittivity of PTSA doped polyanilines and their EMI shielding
performance. Indian Journal of Pure & Applied Physics, 52, 175–182.
39. Song, Y., Wang, H., Zheng, Y., & Xu, C. (2002). Preparation of highly conducting polyaniline
films and study on their electromagnetic shielding properties. Gaofenzi Xuebao, 92–95.
40. Maekelae, T., Pienimaa, S., Taka, T., Jussila, S., & Isotalo, H. (1997). Thin polyaniline films
in EMI shielding. Synthetic Metals, 85, 1335–1336.
41. Ni, X.  W., Hu, S.  X., Zhou, J.  Y., Sun, C.  H., Bai, X.  X., & Chen, P. (2011). Synthesis
and microwave absorbing properties of poly(3,4-ethylenedioxythiophene)(PEDOT) micro-
spheres. Polymers for Advanced Technologies, 22, 532–537.
42. Lee JY, Young K, Han K, Kim MS, Joo JS, Jeong SH, Kim SH, & Byun SW (2002).
Electromagnetic interference shielding fabrics coated with electrically conducting polymers,
PMSE-188.
208 S. K. Srivastava

43. Kim, B. R., Lee, H. K., Kim, E., & Lee, S.-H. (2010). Intrinsic electromagnetic radiation
shielding/absorbing characteristics of polyaniline-coated transparent thin films. Synthetic
Metals, 160, 1838–1842.
44. Niu, Y. (2008). Electromagnetic interference shielding with polyaniline nanofibers composite
coatings. Polymer Engineering and Science, 48, 355–359.
45. Kim, B.  R., Lee, H.  K., Park, S.  H., & Kim, H.  K. (2011). Electromagnetic interference
shielding characteristics and shielding effectiveness of polyaniline-coated films. Thin Solid
Films, 519, 3492–3496.
46. Dhawan, S. K., Singh, N., & Venkatachalam, S. (2001). Shielding effectiveness of conducting
polyaniline coated fabrics at 101 GHz. Synthetic Metals, 125, 389–393.
47. Yu, D., WangY, H. T., & WangW, L. B. (2018). Preparation of silver-plated polyimide fabric
initiated by polyaniline with properties. Journal of Industrial Textiles, 47, 1392–1406.
48. Stempien, Z., Rybicki, T., Rybicki, E., Kozanecki, M., & Szynkowska, M. I. (2015). In-situ
deposition of polyaniline and polypyrrole electroconductive layers on textile surfaces by the
reactive ink-jet printing technique. Synthetic Metals, 202, 49–62.
49. Yang, Z., Munan, Q., Bianying, W., Lele, C., & Ying, Y. (2017). A novel polyaniline-coated
bagasse fiber composite with core–shell heterostructure provides effective electromagnetic
shielding performance. ACS Applied Materials & Interfaces, 9, 809–818.
50. Onar, N., Aksit, A. C., Ebeoglugil, M. F., Birlik, I., Celik, E., & Ozdemir, I. (2009). Structural,
electrical, and electromagnetic properties of cotton fabrics coated with polyaniline and poly-
pyrrole. Journal of Applied Polymer Science, 114, 2003–2010.
51. Unnikrishnan, S. K., Vinayasree, S., Halliah, G. P., & Anantharaman, M. R. (2013). Flexible
electromagnetic interference shields in S band region from textile materials. Journal of
Industrial Textiles, 43, 215–230.
52. Hoghoghifard, S., Mokhtari, H., & Dehghani, S. (2018). Improving EMI shielding effective-
ness and dielectric properties of polyaniline-coated polyester fabric by effective doping and
redoping procedures. Journal of Industrial Textiles, 47, 587–601.
53. Joseph, N., Varghese, J., & Sebastian, M. T. (2017). In situ polymerized polyaniline nanofiber-­
based functional cotton and nylon fabrics as millimeter-wave absorbers. Polymer Journal, 49,
391–399.
54. Joseph, N., Varghese, J., & Sebastian, M.  T. (2016). A facile formulation and excellent
electromagnetic absorption of room temperature curable polyaniline nanofiber based inks.
Journal of Materials Chemistry C, 4, 999–1008.
55. Manesh, F. Y., Hasani, H., & Mortazavi, S. M. (2014). Analyzing the effect of yarn and fab-
rics parameters on electromagnetic shielding of metalized fabrics coated with polyaniline.
Journal of Industrial Textiles, 44, 434–446.
56. Zehra, F., Sagirli, E., & Usta, I. (2014). Electromagnetic shielding effectiveness of polyester
fabrics with polyaniline deposition. Textile Research Journal, 84, 903–912.
57. Tian, M., Du, M., Qu, L., Chen, S., Zhu, S., & Han, G. (2017). Electromagnetic interference
shielding cotton fabrics with high electrical conductivity and electrical heating behavior via
layer-by-layer self-assembly route. RSC Advances, 7, 42641–42652.
58. Wanasinghe, S. D., Aslani, F., Ma, G., & Habib, D. (2020). Review of polymer composites
with diverse nanofillers for electromagnetic interference. Nanomaterials, 10, 541. https://doi.
org/10.3390/nano10030541.
59. Qiao, Y.-S., Shen, L.-Z., Dou, T., & Hu, M. (2010). Polymerization and characterization of
high conductivity and good adhension polypyrrole films for electromagnetic interference
shielding. Chinese Journal of Polymer Science, 28, 923–930.
60. Fauveaux, S., & Miane, J.-L. (2003). Broadband electromagnetic shields using polyaniline
composites. Electromagnetics, 23, 617–627.
61. Hakansson, E., Amiet, A., Nahavandi, S., & Kaynak, A. (2006). Electromagnetic interference
shielding and radiation absorption in thin polypyrrole films. European Polymer Journal, 43,
205–213.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 209

62. Avloni, J., Lau, R., Ouyang, M., Florio, L., Henn, A. R., & Sparavigna, A. (2008). Polypyrrole-­
coated nonwovens for electromagnetic shielding. Journal of Industrial Textiles, 38, 55–68.
63. He, Q., Lv, J., Xu, H., Zhang, L., Zhong, Y., Sui, X., Wang, B., Chen, Z., & Mao, Z. (2019).
Enhancing electrical conductivity and electrical stability of polypyrrole-coated cotton fabrics
via surface microdissolution. Journal of Applied Polymer Science, 136, 47515/1. https://doi.
org/10.1002/app.47515.
64. Marins, J. A., Soares, B. G., Fraga, M., Muller, D., & Barra, G. M. O. (2014). Self-supported
bacterial cellulose polyaniline conducting membrane as electromagnetic interference shield-
ing material: effect of the oxidizing agent. Cellulose, 21, 1409–1418.
65. Gahlout, P., & Choudhary, V. (2017). 5-Sulfoisophthalic acid monolithium salt doped poly-
pyrrole/multiwalled carbon nanotubes composites for EMI shielding application in X-band
(8.2–12.4 GHz). Journal of Applied Polymer Science, 134, 45370. https://doi.org/10.1002/
app.45370.
66. Kim, S.  H., Jang, S.  H., Byun, S.  W., Lee, J.  Y., & Soo, J.  J. Electrical properties and
EMI shielding characteristics of polypyrrole–nylon 6 composite fabrics. Journal of Applied
Polymer Science, 87, 1969–1974. Thermal stability of polypyrrole-nylon 6 composite fabrics
for EMI shielding, PMSE Preprints 86: 173–174.
67. Yildiz, Z., & UstaI, G. A. (2013). Investigation of the electrical properties and electromag-
netic shielding effectiveness of polypyrrole coated cotton yarns. Fibres & Textiles in Eastern
Europe, 21, 32–37.
68. Dandekar, S. S., & Kelkar, D. S. (2008). Application of polypyrrole as an electromagnetic
interference shielder. Indian Journal of Pure and Applied Physics, 46, 215–216.
69. Safarova, V., Gregr, J., & Martinek, M. (2012). Preparation of functional PET fabric/polypyr-
role composite, Scientific papers of the university of Pardubice, series A. Faculty of Chemical
Technology, 18, 117–131.
70. Varshney, S., Ohlan, A., Jain, V.  K., Dutta, V.  P., & Dhawan, S.  K. (2013). Robust multi-
functional free standing polypyrrolesheet for electromagnetic shielding. Science of Advanced
Materials, 5, 1–10.
71. Kim, M. S., Kim, H. K., Byun, S. W., Jeong, S. H., Hong, Y. K., JooJ, S., Song, K. T., Kim,
J. K., Lee, C. J., & Lee, J. Y. (2002). PET fabric/polypyrrole composite with high electrical
conductivity for EMI shielding. Synthetic Metals, 126, 233–239.
72. Gahlout, P., & Choudhary, V. (2019). Microwave shielding behaviour of polypyrrole impreg-
nated fabrics. Composites Part B: Engineering, 175, 107093. https://doi.org/10.1016/j.
compositesb.2019.107093.
73. Abdi, M. M., Kassim, A., Ekramul Mahmud, H. N. M., Yunus, W. M. M., & Talib, Z. A. (2010).
Electromagnetic interference shielding effectiveness of new conducting polymer composite.
Journal of Macromolecular Science, Part A, 47, 71–75.
74. Kim, H.  K., KimMS, C.  S. Y., Park, Y.  H., Jeon, B.  S., Lee, J.  Y., Hong, Y.  K., & JooJ,
K. S. H. (2003). Characteristics of electrically conducting polymer-coated textiles. Molecular
Crystals and Liquid Crystals, 405, 161–169.
75. Wu, F., Sun, M., Jiang, W., Zhang, K., XieA, W.  Y., & Wang, M. (2016). A self-assem-
bly method for the fabrication of a three-dimensional (3D) polypyrrole (­PPy)/poly(3,4-­
ethylenedioxythiophene) (PEDOT) hybrid composite with excellent absorption performance
against electromagnetic pollution. Journal of Materials Chemistry C, 4, 82–88.
76. Dhawan, S. K., Singh, N., & Venkatachalam, S. (2002). Shielding behaviour of conducting
polymer-coated fabrics in X-band, W-band and radio frequency range. Synthetic Metals, 129,
261–267.
77. Yildiz, Z., Usta, I., & Gungor, A. (2012). Electrical properties and electromagnetic shielding
effectiveness of polyester yarns with polypyrrole deposition. Textile Research Journal, 82,
2137–2148.
78. Yamamoto, T., Egami, Y., Nishida, K., Suzuki, K., & Hi, I. (2011). Microwave absorbing
property of stacked polypyrrole-coated nonwoven textiles. Japanese Journal of Applied
Physics, 50, 09NF04-1/6.
210 S. K. Srivastava

79. Hakansson, E., Amiet, A., & Kaynak, A. (2006). Electromagnetic shielding properties of
polypyrrole/polyester composites in the 1–18 GHz frequency range. Synthetic Metals, 156,
917–925.
80. Yuan, Y., Yin, W., Yang, M., Xu, F., Zhao, X., Li, J., Peng, Q., He, X., Du, S., & Li, Y. (2018).
Lightweight, flexible and strong core-shell non-woven fabrics covered by reduced graphene
oxide for high-performance electromagnetic interference shielding. Carbon, 130, 59–68.
81. Lee, C.  Y., Lee, D.  E., Jeong, C.  K., Hong, Y.  K., Shim, J.  H., Joo, J., Kim, M.  S., Lee,
J.  Y., Jeong, S.  H., Byun, S.  W., et  al. (2002). Electromagnetic interference shielding by
using conductive polypyrrole and metal compound coated on fabrics. Polymers for Advanced
Technologies, 13, 577–583.
82. Lee, C. Y., Lee, D. E., Joo, J., Kim, M. S., Lee, J. Y., Jeong, S. H., & Byun, S. W. (2001).
Conductivity and EMI shielding efficiency of polypyrrole and metal compounds coated on
(non) woven fabrics. Synthetic Metals, 119, 429–430.
83. Saini, P., & Choudhary, V. (2013). Electrostatic charge dissipation and electromagnetic inter-
ference shielding response of polyaniline based conducting fabrics. Indian Journal of Pure
and Applied Physics, 51, 112–117.
84. Luo, J., Wang, L., Huang, X., Li, B., Guo, Z., Song, X., Lin, L., Tang, L.-C., Xue, H., & Gao,
J. (2019). Mechanically durable, highly conductive, and anticorrosive composite fabrics with
excellent self-cleaning performance for high-efficiency electromagnetic interference shield-
ing. ACS Applied Materials & Interfaces, 11, 10883–10894.
85. Yu, D., Wang, Y., Hao, T., Wang, W., & Liu, B. (2018). Preparation of silver-plated polyimide
fabric initiated by polyaniline with properties. Journal of Industrial Textiles, 47, 1392–1406.
86. Jang, S. H., Byun, S. W., Kim, S. H., Joo, J. S., Lee, J. Y., & Jeong, S. H. (2002). Preparation
and EMI shielding characteristics of polypyrrole-nylon 6 composite fabrics. Journal of the
Korean Fiber Society, 39, 217–223.
87. Sapurina, I., Kazantseva, N. E., Ryvkina, N. G., Prokes, J., Saha, P., & Stejskal, J. (2005).
Electromagnetic radiation shielding by composites of conducting polymers and wood.
Journal of Applied Polymer Science, 95, 807–814.
88. Fang, F., Li, Y.-Q., Xiao, H.-M., Hu, N., & Fu, S.-Y. (2016). Layer-structured silver nanowire/
polyaniline composite film as a high performance X-band EMI shielding material. Journal of
Materials Chemistry C, 4, 4193–4203.
89. Suna, Y., Guo, G., Yang, B., He, M., Tian, Y., Cheng, J. C., & Liu, Y. (2012). Simple synthesis
of polyaniline microtubes for the application on silver microrods preparation. Journal of
Material Research, 27, 457–462.
90. Chen, Y., LiY, Y.  M., & Tai, N. (2013). Electromagnetic interference shielding efficiency
of polyaniline composites filled with graphene decorated with metallic nanoparticles.
Composites Science and Technology, 80, 80–86.
91. Hu, X.-S., & Shen, Y. (2017). Fabrication of novel polyaniline/flowerlike copper monosulfide
composites with enhanced electromagnetic interference shielding effectiveness. Journal of
Applied Polymer Science, 134, 45232.
92. Panigrahi, R., & Srivastava, S. K. (2015). Trapping of microwave radiation in hollow polypyr-
role microsphere through enhanced internal reflection: A novel approach. Scientific Reports,
5, 7638. https://doi.org/10.1038/srep07638.
93. Mu, S., Xie, H., Wang, W., & Yu, D. (2015). Electroless silver plating on PET fabric initiated
by in situ reduction of polyaniline. Applied Surface Science, 353, 608–614.
94. Wang, Y., Gu, F.-Q., NiLi, J., Liang, K., Marcus, K., Liu, S.-L., Yafng, F., Chen, J.-J., & Feng,
Z.-S. (2017). Easily fabricated and lightweight PPy/PDA/AgNW composites for excellent
electromagnetic interference shielding. Nanoscale, 9, 18318–18325.
95. Lee, C. Y., Song, H. G., Jang, K. S., Oh, E. J., Epstein, A. J., & Joo, J. (1999). Electromagnetic
interference shielding efficiency of polyaniline mixtures and multilayer films. Synthetic
Metals, 102, 1346–1349.
96. Joo, J., Lee, C. Y., Song, H. G., Kim, J.-W., Jang, K. S., Oh, E. J., & Epstein, A. J. (1998).
Enhancement of electromagnetic interference shielding efficiency of polyaniline through mix-
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 211

ture and chemical doping. Molecular Crystals and Liquid Crystals Science and Technology
Section A, 316, 367–370.
97. Bonaldi, R. R., Siores, E., & Shah, T. (2014). Characterization of electromagnetic shielding
fabrics obtained from carbon nanotube composite coatings. Synthetic Metals, 187, 1–8.
98. Erdogan, M. K., Karakisla, M., & Sacak, M. (2018). Polypyrrole and silver particles coated
poly(ethylene terephthalate) nonwoven composite for electromagnetic interference shielding.
Journal of Composite Materials, 52, 1353–1362.
99. Chen, J.-J., Liu, S.-L., Wu, H.-B., Sowade, E., Baumann, R. R., Wang, Y., Gu, F.-Q., Liu,
C.-R.-L., & Feng, Z.-S. (2018). Structural regulation of silver nanowires and their application
in flexible electronic thin films. Materials and Design, 154, 266–274.
100. Zhao, H., Hou, L., & Lu, Y. (2016). Electromagnetic shielding effectiveness and service-
ability of the multilayer structured cuprammonium fabric/polypyrrole/copper (CF/PPy/Cu)
composite. Chemical Engineering Journal, 297, 170–179.
101. Ebrahimi, I., & Gashti, M. P. (2018). Polypyrrole-MWCNT-Ag composites for electromag-
netic shielding: Comparison between chemical deposition and UV-reduction approaches.
Journal of Physics and Chemistry of Solids, 118(2018), 80–87.
102. Bora, P. J., Vinoy, K. J., Ramamurthy, P. C., & Kishore, M. G. (2016). Lightweight polyaniline-­
cobalt coated fly ash cenosphere composite film for electromagnetic interference shielding.
Journal of Composite Materials, 12, 603–609.
103. Hu, J., Li, G., Shi, J., Yang, X., & Xin, D. (2017). Improving the eelectromagnetic shielding
of nickel/polyaniline coated polytrimethylene—Tetraphathalate knitted fabric by optimizing
the electrolass plating conditions. Textile Research Journal, 87, 902–912.
104. Bora, P.  J., Vinoy, K.  J., Ramamurthy, P.  C., & Kishore, M.  G. (2015). Nickel coated fly
ash cenosphere (Ni–FAC) doped polyaniline composite film for electromagnetic shielding.
Materials Research Express, 2, 036403. https://doi.org/10.1088/2053-­1591/2/3/036403.
105. Zhao, H., Hou, L., & Lu, Y. (2016). Electromagnetic interference shielding of layered linen
fabric/polypyrrole/nickel (LF/PPy/Ni) composites. Materials & Design, 95(5), 97–106.
106. Ji, K., Zhao, H., Huang, Z., Dai, Z. (2014). Performance of open-cell foam of Cu–Ni alloy
integrated with graphene as a shield against electromagnetic interference. Mater Lett 122,
244–247.
107. Wang, H., Ma, N., Yan, Z., Deng, L., He, J., Hou, Y., et al. (2015). Cobalt/polypyrrole nano-
composites with controllable electromagnetic properties. Nanoscale, 7, 7189–7196.
108. Qiongzhen, L., Cong, Yi., Jiahui, C., Ming, X., Ying L. et al. (2021). Flexible, breathable,
and highly environmental-stable Ni/PPy/PET conductive fabrics for efficient electromag-
netic interference shielding and wearable textile antennas, Composites Part B: Engineering,
215, 108752.
109. Zhao, H., Hou, L., Bi, S., & Lu, Y. (2017). Enhanced X-band electromagnetic-interference
shielding performance of layer-structured fabric-supported polyaniline/cobalt-nickel coat-
ings. ACS Applied Materials & Interfaces, 9, 33059–33070.
110. Kamchi, N.  E., Belaabed, B., Wojkiewicz, J.-L., Lamouri, S., & Lasri, T. (2013). Hybrid
polyaniline/nanomagnetic particles composites: High performance materials for EMI shield-
ing. Journal of Applied Polymer Science, 127, 4426–4432.
111. Sankaran, S., Deshmukh, K., Ahamed, M. B., & Pasha, S. K. K. (2018). Recent advances in
electromagnetic interference shielding properties of metal and carbon filler reinforced flex-
ible polymer composites: A review. Composites Part A: Applied Science and Manufacturing,
114, 49–71.
112. Lyu, L., Liu, J., Liu, H., Liu, C., Lu, Y., et al. (2018). An overview of electrically conductive
polymer nanocomposites toward electromagnetic interference shielding. Engineered Science,
2, 26–42.
113. Mokry, G., de Nicolás, M., Baselga, J., & González, J.  P. (2016). Carbon nanotube com-
posites as electromagnetic shielding materials in GHz range. In M. R. Berber & I. H. Hafez
(Eds.), Carbon nanotubes-current progress of their polymer composites. Intech Open. https://
doi.org/10.5772/62508.
212 S. K. Srivastava

114. Mostaani, F., Moghbeli, M. R., & Karimian, H. (2018). Electrical conductivity, aging behav-
ior, and electromagnetic properties of polyaniline/MWCNT nanocomposites. Journal of
Thermoplastic Composite Materials, 31, 1393–1415.
115. Mishra, M., Singh, A.  P., Gupta, V., Chandra, A., & Dhawan, S.  K. (2016). Tunable EMI
shielding effectiveness using new exotic carbon: Polymer composites. Journal of Alloys and
Compounds, 688, 399–403.
116. Makeiff, D. A., & Huber, T. (2006). Microwave absorption by polyaniline-carbon nanotube
composites. Synthetic Metals, 156, 497–505.
117. Saini, P., Choudhary, V., Singh, B. P., et al. (2009). Polyaniline-MWCNT nanocomposites for
microwave absorption and EMI shielding. Materials Chemistry and Physics, 113, 919–926.
118. Sharma, A. K., Bhardwaj, P., Singh, K. K., & Dhawan, S. K. (n.d.). Improved microwave
shielding properties of polyaniline grown over three-dimensional hybrid carbon assemblage
substrate. Applied Nanoscience, 5, 635–644.
119. Jelmy, E. J., Ramakrishnan, S., & Kothurkar, N. K. (2016). EMI shielding and microwave
absorption behavior of Au-MWCNT/polyaniline nanocomposites. Polymers for Advanced
Technologies, 27, 1246–1257.
120. Gupta, T.  K., Singh, B.  P., Mathur, R.  B., & Dhakate, S.  R. (2014). Multi-walled carbon
nanotube–graphene–polyaniline multiphase nanocomposite with superior electromagnetic
shielding effectiveness. Nanoscale, 6, 842–851.
121. Yun, J., & Kim, H.-I. (2012). Electromagnetic interference shielding effects of polyaniline-­
coated multi-wall carbon nanotubes/maghemitenanocomposites. Polymer Bulletin, 68,
561–573.
122. Kim, Y. Y., Yun, J., Kim, H.-I., & Lee, Y.-S. (2011). Effect of oxyfluorination on electromag-
netic interference shielding of polyaniline-coated multi-walled carbon nanotubes. Colloid &
Polymer Science, 289, 1749–1755.
123. Cao, M.-S., Yang, J., Song, W.-L., Zhang, D.-Q., Wen, B., Jin, H.-B., Hou, Z.-H., & Yuan,
J. (2012). Ferroferric oxide/multiwalled carbon nanotube vs polyaniline/ferroferric oxide/
multiwalled carbon nanotube multiheterostructures for highly effective microwave absorp-
tion. ACS Applied Materials & Interfaces, 4, 6949–6956.
124. Kausar, A., Meer, S., & Iqbal, T. (2017). Structure, morphology, thermal, and electro-­
magnetic shielding properties of polystyrene microsphere/polyaniline/multi-walled carbon
nanotube nanocomposite. Journal of Plastic Film and Sheeting, 33, 262–289.
125. Saini, P., Choudhary, V., Singh, B.  P., Mathur, R.  B., & Dhawan, S.  K. (2011). Enhanced
microwave absorption behavior of polyaniline-CNT/polystyrene blend in 12.4–18.0  GHz
range. Synthetic Metals, 161, 1522–1526.
126. Bingqing, Y., Liming, Y., Leimei, S., Kang, A., & Xinluo, Z. (2012). Comparison of elec-
tromagnetic interference shielding properties between single-wall carbon nanotube and gra-
phene sheet/polyaniline composites. Journal of Physics D: Applied Physics, 45, 235108.
127. Khasim, S. (2019). Polyaniline-Graphene nanoplatelet composite films with improved con-
ductivity for high performance X-band microwave shielding applications. Results in Physics,
12, 1073–1081.
128. Yu, H., Wang, T., Wen, B., Lu, M., Xu, Z., Zhu, C., Chen, Y., Xue, X., Sun, C., Cao, M.,
Yu, H., Wang, T., Wen, B., Lu, M., Xu, Z., et  al. (2012). Graphene/polyaniline nanorod
arrays: Synthesis and excellent electromagnetic absorption properties. Journal of Materials
Chemistry, 22, 21679–21685.
129. Yun, J., Im, J. S., & Kim, H.-I. (2012). Effect of oxygen plasma treatment of carbon nano-
tubes on electromagnetic interference shielding of polypyrrole-coated carbon nanotubes.
Journal of Applied Polymer Science, 126, E39–E47.
130. Kim, Y.-Y., Yun, J., Kim, H.-I., & Lee, Y.-S. (2012). Effect of oxyfluorination on electromag-
netic interference shielding of polypyrrole-coated multiwalled carbon nanotubes. Journal of
Industrial and Engineering Chemistry, 18, 392–398.
131. Mohan, R.  R., Varma, S.  J., Faisal, M., & Jayalekshmi, S. (2015). Polyaniline/graphene
hybrid as an effective broadband electromagnetic shield. RSC Advances, 5, 5917–5923.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 213

132. Modak, P., Kondawar, S. B., & Nandanwar, D. V. (2015). Synthesis and characterization of
conducting polyaniline/graphene nanocomposites for electromagnetic interference shielding.
Procedia Materials Science, 10, 588–594.
133. Zhang, W., Zhang, X., Qiao, Y., Yan, H., & Qi, S. (2018). Covalently bonded GNPs-NH-­
PANI nanorod arrays modified by Fe3O4 nanoparticles as high-performance electromagnetic
wave absorption materials. Materials Letters, 216, 101–105.
134. Kang, S., Qiao, S., Hu, Z., Yu, J., Wang, Y., & Zhu, J. (2019). Interfacial polymerized reduced
graphene oxide covalently grafted polyaniline nanocomposites for high-performance electro-
magnetic wave absorber. Journal of Materials Science, 54, 6410–6424.
135. Hu, X.-S., Shen, Y., Lu, L.-S., Xu, J., & Zhen, J.-J. (2017). Enhanced electromagnetic inter-
ference shielding effectiveness of ternary PANI/CuS/RGO composites. Journal of Materials
Science: Materials in Electronics, 28, 6865–6872.
136. Wang, C., Ding, Y., Yuan, Y., He, X., Wu, S., Hu, S., Zou, M., Zhao, W., Yang, L., Cao, A.,
et al. (2015). Graphene aerogel composites derived from recycled cigarette filters for electro-
magnetic wave absorption. Journal of Materials Chemistry C, 3, 11893–11901.
137. Bhardwaj, P., Kaushik, S., Gairola, P., & Gairola, S. P. (2018). Exceptional electromagnetic
radiation shielding performance and dielectric properties of surfactant assisted polypyrrole-
-carbon allotropes composites. Radiation Physics and Chemistry, 151, 156–163.
138. Wu, F., Xie, A., Sun, M., Wang, Y., & Wang, M. (2015). Reduced graphene oxide (RGO)
modified spongelikepolypyrrole (PPy) aerogel for excellent electromagnetic absorption.
Journal of Materials Chemistry A, 3, 14358–14369.
139. Zhang, J., Shi, C., Ji, T., Wu, G., & Kou, K. (2014). Preparation and microwave absorb-
ing characteristics of multi-walled carbon nanotube/Chiral-polyaniline composites. Open
Journal of Polymer Chemistry, 4, 62–72. http://www.scirp.org/journal/ojpchem.
140. Saini, P., Arora, M., Gupta, G., Gupta, B. K., Singh, V. N., & Choudhary, V. (2013). High
permittivity polyaniline–barium titanate nanocomposites with excellent electromagnetic
interference shielding response. Nanoscale, 5, 4330–4336.
141. Gonzalez, M., Soares, B.  G., Magioli, M., Marins, J.  A., & Rieumont, J. (2012). Facile
method for synthesis of polyaniline/silica hybrid composites by simultaneous sol–gel pro-
cess and “in situ” polymerization of aniline. Journal of Sol-Gel Science and Technology, 63,
373–381.
142. Zhang, D., Cheng, J., Yang, X., Zhao, B., & Cao, M. (2014). Electromagnetic and microwave
absorbing properties of magnetite nanoparticles decorated carbon nanotubes/polyaniline
multiphase heterostructures. Journal of Materials Science, 49, 7221–7230.
143. Varshney, S., Ohlan, A., Jain, V. K., Dutta, V. P., & Dhawan, S. K. (2014). Synthesis of ferro-
fluid based nano architecture polypyrrole composites and its application for electromagnetic
shielding. Materials Chemistry and Physics, 143, 806–813.
144. Singh, K., Ohlan, A., Pham, V. H., Balasubramaniyan, R., Varshney, S., Jang, J., Hur, S. H.,
Choi, W.  M., Kumar, M., Dhawan, S.  K., et  al. (2013). Nanostructured graphene/Fe3O4
incorporated polyaniline as a high performance shield against electromagnetic pollution.
Nanoscale, 5, 2411–2420.
145. Bera, R., Das, A. K., Maitra, A., Paria, S., Karan, S. K., & Khatua, B. B. (2017). Salt leached
viable porous Fe3O4 decorated polyaniline-SWCNH/PVDF composite spectacles as an admi-
rable electromagnetic shielding efficiency in extended Ku-band region. Composites Part B:
Engineering, 129, 210–220.
146. Bera, R., Paria, S., Karan, S. K., Das, A. K., Maitra, A., & Khatua, B. B. (2017). NaCl leached
sustainable porous flexible Fe2O3 decorated RGO-polyaniline/PVDF composite for durable
application against electromagnetic pollution. Express Polymer Letters, 11, 419–433.
147. Saini, P., Choudhary, V., Vijayan, N., & Kotnala, R.  K. (2012). Improved electromagnetic
interference shielding response of poly(aniline)-coated fabrics containing dielectric and mag-
netic nanoparticles. Journal of Physical Chemistry C, 116, 13403–13412.
148. Saini, P., Barala, S. K., Arora, M., & Kotnala, R. K. (2013). Designing of conducting polymer
composites for shielding of microwave radiations. In AIP Conference Proceedings 1536
214 S. K. Srivastava

Proceeding of International Conference on Recent Trends in Applied Physics & Material


Science (pp. 1237–1239).
149. Moon, Y.-E., Yun, J., & Kim, H.-I. (2013). Synergetic improvement in electromagnetic inter-
ference shielding characteristics of polyaniline-coated graphite oxide/γ-Fe2O3/BaTiO3 nano-
composites. Journal of Industrial and Engineering Chemistry, 19, 493–497.
150. Singh, K., Ohlan, A., Bakhshi, A.  K., & Dhawan, S.  K. (2010). Synthesis of conducting
ferromagnetic nanocomposite with improved microwave absorption properties. Materials
Chemistry and Physics, 119, 201–207.
151. Liu, P., Huang, Y., Yang, Y., Yan, J., & Zhang, X. (2016). Sandwich structures of graphene@
Fe3O4@PANI decorated with TiO2 nanosheets for enhanced electromagnetic wave absorption
properties. Journal of Alloys and Compounds, 662, 63–68.
152. Dorraji, M.  S. S., Rasoulifard, M.  H., Khodabandeloo, M.  H., Rastgouy-Houjaghan, M.,
& Zarajabad, H.  K. (2016). Microwave absorption properties of polyaniline-Fe3O4/ZnO-­
polyester nanocomposite: Preparation and optimization. Applied Surface Science, 366,
210–218.
153. Liu, P., Huang, Y., Wang, L., & Zhang, W. (2013). Synthesis and excellent electromag-
netic absorption properties of polypyrrole-reduced graphene oxide-Co3O4 nanocomposites.
Journal of Alloys and Compounds, 573, 151–156.
154. Belaabed, B., Wojkiewicz, J. L., Lamouri, S., Kamchi, N. E., & Lasri, T. (2012). Synthesis
and characterization of hybrid conducting composites based on polyaniline/magnetite fillers
with improved microwave absorption properties. Journal of Alloys and Compounds, 527,
137–144.
155. Gupta, A., Varshney, S., Goyal, A., Sambyal, P., Gupta, B.  P., & Dhawan, S.  K. (2015).
Enhanced electromagnetic shielding behaviour of multilayer graphene anchored luminescent
TiO2 in PPY matrix. Materials Letters, 158, 167–169.
156. Olad, A., & Shakoori, S. (2018). Electromagnetic interference attenuation and shielding effect
of quaternary nanocomposite as a broad band microwave-absorber. Journal of Magnetism
and Magnetic Materials, 458, 335–345.
157. Zhang, C., Chen, Y., Li, H., Tian, R., & Liu, H. (2018). Facile fabrication of three-­dimensional
lightweight RGO/PPy nanotube/Fe3O4 aerogel with excellent electromagnetic wave absorp-
tion properties. ACS Omega, 3, 5735–5743.
158. Wei, S., Wang, Q., Zhu, J., Sun, L., Lin, H., & Guo, Z. (2011). Multifunctional composite
core–shell nanoparticles. Nanoscale, 3, 4474–4502.
159. Azadmanjiri, J., Suzuki, K., Selomulya, C., Amiet, A., Cashion, J. D., & Simon, G. P. (2012).
The use of plasma treatment for simultaneous carbonization and reduction of iron oxide/
polypyrrole core/shell nanoparticles. Journal of Nanoparticle Research, 14, 1078/1–1078/11.
160. Wu, Z., Tan, D., Tian, K., Hu, W., Wang, J., Su, M., & Li, L. (2017). Facile preparation of
core-shell Fe3O4@Polypyrrole composites with superior electromagnetic wave absorption
properties. Journal of Physical Chemistry C, 121, 15784–15792.
161. Manna, M., & Srivastava, S. K. (2017). Fe3O4@Carbon@Polyaniline trilaminar core−shell
composites as superior microwave absorber in shielding of electromagnetic pollution. ACS
Sustainable Chemistry & Engineering, 5, 10710–10721.
162. Manna, K., & Srivastava, S. K. (2020). Tunning of shells in Trilaminar Core@Shell nano-
composites in controlling electromagnetic interference through switching of the shielding
mechanism. Langmuir, 36(16), 4519–4531.
163. Movassagh-Alanagh, F., Bordbar-Khiabani, A., & Ahangari-Asl, A. (2017). Three-phase
PANI@nano-Fe3O4@CFs heterostructure: Fabrication, characterization and investigation of
microwave absorption and EMI shielding of PANI@nano-Fe3O4@CFs/epoxy hybrid com-
posite. Composites Science and Technology, 150, 65–78.
164. Wang, L., Huang, Y., Li, C., Chen, J., & Sun, X. (2014). Enhanced microwave absorption
properties of N-doped graphene@PANInanorod arrays hierarchical structures modified by
Fe3O4 nanoclusters. Synthetic Metals, 198, 300–307.
7  Intrinsically Conducting Polymer Nanocomposites in Shielding of Electromagnetic… 215

165. Bhattacharjee, Y., Chatterjee, D., & Bose, S. (2018). Core-multishell heterostructure with
excellent heat dissipation for electromagnetic interference shielding. ACS Applied Materials
& Interfaces, 10, 30762–30773.
166. Xu, P., Han, X., Wang, C., Zhou, D., Lv, Z., Wen, A., et  al. (2008). Synthesis of electro-
magnetic functionalized nickel/Polypyrrole core/shell composites. The Journal of Physical
Chemistry B, 112, 10443–10448.
167. Dong, X. L., Zhang, X. F., Huang, H., & Zuo, F. (2008). Enhanced microwave absorption in
Ni/polyaniline nanocomposites by dual dielectric relaxations. Applied Physics Letters, 92,
013127. https://doi.org/10.1063/1.2830995.
168. Li, C., Ji, S., Jiang, X., Waterhouse, G. I. N., Zhang, Z., & Yu, L. (2018). Microwave absorp-
tion by watermelon-like microspheres composed of γ-Fe2O3, microporous silica and polypyr-
role. Journal of Materials Science, 53, 9635–9649.
169. Liu, P., Huang, Y., Yana, J., & Zhaob, Y. (2016). Magnetic graphene@PANI@porous TiO2
ternary composites for high-performance electromagnetic wave absorption. Journal of
Materials Chemistry C, 4, 6362–6370.
Chapter 8
Nanostructuring of Hybrid Materials
Using Wrapping Approach to Enhance
the Efficiency of Visible Light-Responsive
Semiconductor Photocatalyst

V. Vinesh, A. R. Mahammed Shaheer, and B. Neppolian

8.1  Introduction

In recent years, the semiconductor photocatalysis is an economicaly viable method


that plays an important role in the degradation of toxic organic pollutants including
antibiotics, textile dyes, insectisides, drinking water purification, removal of heavy
metals and disinfection. TiO2 is observed to be a promising material due to its chem-
ical and biological stability, water insolubility, and non-toxic, good optical proper-
ties and it can be easily attached with different supporting materials. Since TiO2 has
a bandwidth energy of 3.2 eV, it can only be excited through the ultraviolet (UV)
region (λ < 390 nm), which unfortunately limits its applications [1, 2]. Further, many
methods have been adopted for improvement of the perfomance of TiO2, such as
doping of metals incorporating with low bandgap semiconductor, using supporting
materials etc. Because of the interesting electrical and mechanical properties, the
high surface area, graphene is used as a supporting material for TiO2 in the form
of different nanostructured hybrid materials. Furthermore, graphene can hybridize
with metals, polymers, and metal oxides that have been developed for various
other applications. In the formation of nanohybrids, it is important to consider the
growth of nanocrystasl in the graphene sheet [3–5]. it is noticed that  controlled
nucleation and growth allow chemical interactions and optimal connections between
nanocrystals and graphene sheets, which lead to a very strong electrical and mechan-
ical coupling within the hybrid [6–8].
In addtion, various methodologies have been established to improve the charge
carrier separation of metal oxide photocatalysts, which is yet another problem to be
circumvented. For example, the synthesis of mixed metal oxides and the loading of
noble metals as co-catalysts have proved to be effective methods to overcome the

V. Vinesh · A. R. M. Shaheer · B. Neppolian (*)


Energy and Environmental Remediation Lab, SRM Research Institute, SRM Institute of
Science and Technology, Kattankulathur, Chennai, Tamil Nadu, India
e-mail: neppolib@srmist.edu.in

© Springer Nature Switzerland AG 2021 217


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_8
218 V. Vinesh et al.

issue. On the other hand, the addition of graphene as a solid support to the metal
oxide photocatalysts can also improve the separation of the charge carrier effectively
[9–13] as discused eartlier. Photogenerated electrons can rapidly migrate to the gra-
phene sheet because graphene has a high electronic conductivity, which improves
the overall  photocatalytic activity [14–16]. Thus, in this chapter, we describe the
synthesis of the semiconductor photocatalysts that responds to visible light through
hybrid materials using a wrapping up approach [17–19]. The main focus is to develop
different semiconductors wrapped  sphotocatalysts with supports  for the improve-
ment of the degradation of pollutants and H2 production applications [20–25].

8.2  S
 ynthesis of Nanostructuring of Hybrid Materials Using
Wrapping Approach

8.2.1  P
 reparation of Graphene Oxide Encapsulated TiO2
Core/Shell Microspheres

TiO2 core/shell microspheres encapsulated by graphene oxide were prepared via


three-step process (Fig. 8.1). At the first, modified Hummers method from graphite
powder was adopted to synthesis graphite oxide. At the second, controlled hydroly-
sis of TBOT (tetrabutyl titanate) in ethanol ledad to the formation of monodispersed
spherical TiO2 particles. In detail, to 1.2 mL of KCl aqueous solution, 300 mL of
ethanol was mixed, further followed by 5.4 mL TBOT under constant stirring. After
stirring, the products were centrifuged and washed for several times with distilled
water and ethanol [1]. Then the product kept drying at 60 °C for 12 h. At the end,

Fig. 8.1  Illustration of the preparation procedure of graphene oxide encapsulated TiO2@GO core/
shell microspheres. {Reproduced with permission from ref. [1], Copyright (2013), Elsevier}
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 219

TiO2 core/shell microspheres encapsulated by graphene oxide were fabricated by a


simple and pure chemical bonding reactions involved with dehydration condensa-
tion without any modified agents. Typically, required amount of prepared TiO2
microspheres and graphene oxide mixture was added to 40 mL of distilled water on
continuous stirring, respectively, and then further stirred for 20 h for the encapsula-
tion of graphene oxide. After the reaction, the product was collected and centrifuged
and washed with distilled water for three times and dried at 60 °C overnight. At last
the sample was calcinated at 500 °C for 2.0 h [2–8].

8.2.2  Synthesis of GP Strongly Wrapped TiO2 Photocatalyst

First 0.2 g of TiO2 was dispersed in 100 mL of ethanol by sonication and magnetic
stirring (Fig. 8.2). Then, to form a homogeneous suspension, APTMS (1 mL) was
added followed by heating and reflux at 85 °C for 2 h. TiO2 treated with APTMS

Fig. 8.2  Schematic illustration of synthesis processes for GP (TiO2-graphene) strongly wrapped
and GP-supported anatase TiO2 nanosheets with exposed {001} facets. With SEM images.
{Reproduced with permission from ref. [9], Copyright (2014), Elsevier}
220 V. Vinesh et al.

was rinsed with ethanol to wash the remaining APTMS and dried adequately [9]. A
different amount of negatively charged (graphene oxide) GO suspension (0.5 mg/
mL) was added to the TiO2 dispersion functionalized with positively charged aque-
ous amines (20 mL) to modify the weight ratio between GO and TiO2 (0:1, 0.005:1,
0.01:1, 0.02:1, and 0.05:1) under vigorous stirring. After mixing for 15  min, the
mixture was centrifuged and washed using deionized water. However, for the reduc-
tion of GO in GP, the hydrothermal process was adopted as follows. The TiO2
wrapped with GO (0.2 g) was dispersed in the mixture of deionized water (20 mL)
and ethanol (20 mL), followed by hydrothermal treatment in a Teflon-coated stain-
less steel vessel at 120 ° C for 12 h. After the hydrothermal reaction, the product was
centrifuged and dried at 60 °C overnight [10, 11].

8.2.3  Fabrication of rGO-Wrapped TiO2 Nanofibers (TNFs)

Common modified Hummers method was  adopted for the synthesis of graphite
oxide [26, 27] (Fig. 8.3). For the preparation of the TNF / rGO compounds, 1.25 mg,
2 mg, and 5 mg of graphite oxide were added in 60 mL of deionized water in sonica-
tion for 1  h, to obtain a homogeneous dispersion of GO [11–14]. Then 5  mg of
TNFs was added to each of these GO solutions with constant stirring in the dark, at
room temperature to obtain homogeneous mixtures. The water was slowly evapo-
rated from these mixtures by hot plate drying with continuous stirring in the dark.
Finally, the brown samples were collected and annealed at 400 °C for 2 h. in an
argon atmosphere to reduce the GO sheets to form TNFs wrapped with reduced
graphene oxide (rGO) at ratios of 1: 0.25, 1: 0.5, and 1: 1 for TiO2: GO.

Fig. 8.3  Schematics of reduced graphene oxide-wrapped anatase mesoporous TiO2 nanofibers.
{Reproduced with permission from ref. [13], Copyright (2014), Elsevier}
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 221

8.2.4  Synthesis of GO/Bi2WO6 Composites

Modified Hummers method was adopted using sodium nitrate, potassium perman-


ganate, and sulfuric acid in synthesis of GO (Fig. 8.4). Further, modified hydrother-
mal method was used for uniform growth of Bi2WO6 hierarchical microspheres. In
detail, in 40 mL of deionized water, (1.25 mmol) Na2WO4·2H2O was dissolved. On
the other hand, 40 mL of ethylene glycol (2.5 mmol) Bi(NO3)3 5H2O was dissolved.
Now, the two solutions were mixed and kept stirring for 2 h. Then the suspension
was sealed in a Teflon-lined stainless-steel autoclave of 100 mL capacity, kept at
180 °C for 24 h. Finally, the product was cooled and centrifuged and washed with
DI water and dried at 40 °C in an oven. Moreover, followed by the synthesis of GO/
Bi2WO6 composites, GO aqueous suspension (1.36%) was diluted in deionized
water and dispersed by sonication for 1 h to achieve a brown homogeneous aqueous
suspension. Then 0.1 g of Bi2WO6 microspheres were dispersed in deionized water
by sonication for 30 min. The obtained solution was mixed to the GO suspension
followed by continuous sonication for 2 h. The resulting suspension was frozen by
liquid nitrogen and freeze-drying dehydration to obtain the GO/Bi2WO6 composites
[18–20].

Fig. 8.4  Structure of GO/Bi2WO6 composites—(a) SEM image of synthesized Bi2WO6 micro-
spheres; (b) TEM image of the single Bi2WO6 microspheres; (c) TEM image of the edge of
Bi2WO6; (d) SEM image of GO/Bi2WO6 composites microspheres; (e) TEM image of the single
GO/Bi2WO6 composites microspheres; (f) HRTEM image of GO/ Bi2WO6 composites.
{Reproduced with permission from ref. [19], Copyright (2015), Elsevier}
222 V. Vinesh et al.

8.2.5  Synthesis of W2C@C/HTMs Heterojunction

The schematic representation of W2C@C/HTMs heterojunction is depicted in


(Fig. 8.5a). A simple wet impregnation method was adopted to synthesis of W2C@C/
HTMs heterojunction. In this process, W2C@C was prepared by simple calcination
of dicyanodiamide and ((NH4)5H5[H2(WO4)6]·H2O), and hollow TiO2 microspheres
(HTMs) were prepared by combination of hydrothermal and combustion methods.
Above W2C@C and HTMs were dispersed separately in 25 mL of deionized water
by ultrasonication. Further, prepared solution mixed together under constant stirring
and maintained at 80 °C for 5 h and product collected by precipitation at 100 °C. The
resultant product was annealed at 350 °C for 2 h in Ar atmosphere. In addition, the
thick layer of carbon wrapping on W2C is clearly shown in the TEM images in
(Fig. 8.5b, c) [24].

8.2.6  S
 ynthesis of Wrinkled Graphene-Wrapped
TiO2 Nanotubes

A  schematic representation of the synthesis of wrinkled graphene-wrapped TiO2


nanotubes is shown in (Fig. 8.6a). Graphene oxide (GO) synthesized by modified
hummer method was dispersed in the aqueous solution and, then, above dispersed
GO added to TiO2 suspension and stirred for 1 h. The TiO2 oxygen group bind with
carboxylate functional group of the GO makes dispersed particle to shuttle down in
the suspension. Thereafter, the suspension transferred to hydrothermal reactor and

Fig. 8.5 (a) Schematic
illustration of the synthesis
of W2C@C/HTMs
heterojunction and (b, c)
TEM images of W2C@C/
HTMs heterojunction
[Copyright 24]
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 223

Fig. 8.6  A schematic representation of the synthesis of wrinkled graphene-wrapped TiO2 nano-
tubes (a), SEM images of (b) wrinkled GO and (c & d) wrinkled graphene-wrapped TiO2 nano-
tubes, respectively [Copyright 25]

heated at 130 °C. The GOs functional groups were reduced to form reduced gra-


phene oxide (rGO) by thermal process and chemically Ti–O–C bond formed
between the TiO2 nanotubes and rGO. Further, SEM images GO wrinkles and wrin-
kled graphene-wrapped TiO2 nanotubes are illustrated in (Fig. 8.6b–d). Which con-
firm the complete coverage of rGO wrapping on TiO2 nanotubes i.e. the formation
of wrinkled graphene-wrapped TiO2 nanotubes [25].

8.3  Photocatalytic Application

8.3.1  Photocatalytic Degradation of Organic Pollutants

The metal oxides semiconductors photocatalysis has gained abundant attention in


energy conversion and environmental remediation applications. However, the rapid
recombination of pairs of photogenerated electron and holes suppresses the photo-
catalytic activity of photocatalysts and limits their applications. In addition to con-
ventional methods, such as the doping of noble metals and the making of mixed
metal oxides, it has also been reported that the addition of rGO matrix to metal
oxide photocatalysts is a simple method to improve the rate of charge carrier separa-
tion. In this section, we will analyze and summarize the various nanostructures of
hybrid materials using a wrapping approach to improve the efficiency of photocata-
lytic semiconductors sensitive to visible light on degradation of organic pollutants.
224 V. Vinesh et al.

Hui Liu et al. synthesized TiO2 encapsulated by graphene oxide core/shell struc-
tures which showed an enhanced photocatalytic activity. Moreover, the encapsu-
lated process of this study was mainly based on the condensation reaction of direct
dehydration between the oxygen-containing functional groups (–COOH) on the
surface of graphene oxides and hydroxyl groups (–OH) in TiO2 microspheres sur-
face. Here the graphene oxide was  combined with  TiO2 microspheres to form a
core/shell structure by pure chemical bonding reactions as dehydration condensa-
tion between the hydroxyl groups (AOH) on TiO2 microspheres surface and oxygen-­
containing functional groups (ACOOH) on the surface of graphene oxides. Moreover
this composite exhibited a red shift of the band-edge about 425 nm, which enhanced
photocatalytic and excellent recyclability property under both UV and visible light
for the degradation of Rh B. Here, the photodegradation time correspond to 25 min
in UV irradiation and 130 min in visible light irradiation are shown in (Fig. 8.7d),
which indicated that the sample showed an excellent photocatalytic activity for the
degradation of Rh B [1] (Fig. 8.7a–d).
Yaru Ni, Wei Wang et al. successfully synthesized three-dimensional photoactive
GP strongly wrapped in TiO2 dominated by highly reactive facets {001} using sur-
face modification and the hydrothermal method. The TiO2 strongly wrapped by the
GP showed a strong red shift of the edge of the band and a significant reduction of

Fig. 8.7 (a) UV–vis diffuse reflection spectra of the GO, anatase TiO2, and TiO2@GO core/shell
microspheres. (b and c) Absorption changes of RhB in the presence of TiO2@GO core/shell micro-
spheres under UV and visible irradiation, respectively. (d) Comparison C/C0 plot of photocatalytic
activities of samples TiO2 microspheres, TiO2@GO core/shell microspheres, and P25 {Reproduced
with permission from ref. [1], Copyright (2013), Elsevier}
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 225

the band gap compared to the random TiO2 GP when the same amount of GP was
introduced. GP-wrapped TiO2 prepared as such has excellent photocatalytic proper-
ties under the xenon lamp and visible light irradiation for the  MB degradation,
which is much higher than the commercial P25. Also the GP content influences the
photocatalytic activity of TiO2 with GP envelope. Furthermore, the mechanism of
photocatalytic reactions in Methylene blue (MB) degradation  was proposed. The
strategy presented in this study will provide new perspectives in the preparation of
new highly reactive GP-based photocatalysts and sufficient use of GPs for commer-
cial applications [9] (Fig. 8.8).
Naoki Fukata et al. reported that rGO-wrapped anaerobic TNFs were prepared
using the electrospinning technique along with simple chemical methods. The
advantage of this method is that it does not require solvents or toxic chemicals to
reduce graphene oxide. Wrapping with rGO leads to an efficient photogenerated
charge carrier separation across the interface of rGO and TNFs. The rGO wrapped
with TiO2 by a simple hot plate drying could prevent the agglomeration of the com-
posite. The fibrous morphology was clearly evident from the SEM observations
(Fig. 8.9). HR-TEM images and EELS spectroscopy confirmed that the TNF sur-
face was surrounded by rGO. N2 adsorption-desorption isotherms demonstrated the
mesoporous nature of NF. For the compound, a photocatalytic degradation of 96%
methyl orange (MO) was obtained compared to the degradation of only 43%

Fig. 8.8  Proposed mechanisms for the photocatalytic degradation of MB by (a) GP-supported
TiO2 and (b) GP-wrapped TiO2 under the visible light irradiation. (c) Fluorescence spectra changes
with time of visible light-irradiated TG2 in the terephthalic acid solution and (d) comparison of
fluorescence spectra of terephthalic acid solution with the presence of TG2 and TG2(mix).
{Reproduced with permission from ref. [9], Copyright (2014), Elsevier}
226 V. Vinesh et al.

Fig. 8.9 (a) PL spectra of TNF and TNFG. (b) Photodegradation of MO by TNF and TNFG. Inset
of (a) shows the band alignment at the interface of TNFs and rGO. {Reproduced with permission
from ref. [13], Copyright (2014), Elsevier}

with the naked TNF. This simple synthetic method can be applied not only to wrap
rGO over other nanostructures of different morphology but also for enhancing the
properties of multifunctional materials [13] (Fig. 8.9a, b).
J. Zhai et al. successfully developed GO/Bi2WO6 composites via easy approach
without any surface modification on the Bi2WO6 microspheres. GO/Bi2WO6–10%
was the optimized catalyst for the photocatalytic degradation of RhB and was
almost 5 times higher than those of pure Bi2WO6. The wrapping of GO leads to the
via rapid photogenerated electron holes pairs separation which enhanced the photo-
catalytic activity of the GO/Bi2WO6. Furthermore, when 0.1 mol/L KI was added
into the reaction solution, the photocatalytic reaction was strongly suppressed, indi-
cating that photogenerated hole which was responsible for the degradation process.
GO wrapped on the surface of Bi2WO6; the photogenerated holes on Bi2WO6 could
transfer easily to GO via the well-developed interface. It was expected that hydroxyl
radicals were produced in the system with GO/Bi2WO6 composites. Therefore, the
main oxidative species of the GO/Bi2WO6 system would be holes and hydroxyl
radicals. Moreover, XPS analysis that the carbon bonding composition ratios of GO
have no notable differences before and after the photocatalysis, indicating that the
better stability after photocatalytic process for GO/Bi2WO6 composites
(Fig. 8.10a, b) [19].
Neppolian et al. have successfully synthesized rGO-wrapped Cu2O spheres and
loaded over C3N4 photocatalyst. The rGO was placed exactly at the interfacial of
Cu2O and C3N4 heterojunction. However, the Cu2O–rGO–C3N4 composite covered
the entire solar spectrum with a significant absorption intensity. rGO-wrapped Cu2O
loading caused a red shift in the absorption with respect to considering the absorp-
tion of bare C3N4. Interestingly, a synergistic effect of 2 times was obtained with
Cu2O–rGO–C3N4 for the photocatalytic reduction of 4-nitrophenol to 4-­aminophenol
with respect to Cu2O–rGO and C3N4 and reused for 4 time without any loss in the
photocatalytic activity [14].
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 227

Fig. 8.10 (a) the GO/Bi2WO6 composites band structure and charge separation mechanism and
(b) photocatalytic degradation of RhB mechanism for as prepared composites. (c) The photodeg-
radation plots and (d) degradation rate constant k as a function of GO content. Inserts: first-order
kinetics plot for the photodegradation of RhB. {Reproduced with permission from ref. [19],
Copyright (2015), Elsevier}

8.3.2  Photocatalytic Hydrogen Production

Photocatalytic water splitting to produce renewable H2 production on the surface of


a semiconductor photocatalyst has gained greater attention due to its sustainability
and simplicity for the future renewable energy resource [28, 29]. TiO2 is one of the
best photocatalysts for water splitting application owing to its suitable band position
for reduction of water, availability, chemical stability, low toxicity, and cost-­
effectiveness [30, 31]. However, wide band gap and slow reaction kinetics as a
result of fast recombination of photogenerated charge carriers in the TiO2 hinder its
photocatalytic performance [32, 33]. Several strategies were used to overcome the
above problems, such as doping, heterojunction with narrow band gap materials,
co-catalyst, plasmonic metals incorporation, etc. [31–34]. Among them, the forma-
tion of heterojunction with narrow band gap materials showed considerably sup-
pressed recombination of charge carrier effectively [26, 27].
228 V. Vinesh et al.

Recently, carbon-based materials like graphene, carbon nanotubes, carbon dots,


etc. form heterojunction with TiO2 and help in rapid separation of charge carriers
attributed to its electronic conductivity [35–37]. Off late, wrapping of the carbon on
the TiO2 surface formed the effective and uniform interface between them and pre-
vented the rapid recombination of charge carriers [24, 25]. Yue et al. constructed
W2C@C core/shell by simple calcination method and further W2C@C wrapped on
the hollow TiO2 microspheres (HTMs) by simple wet impregnation method. The
prepared W2C@C/HTMs heterojunctions were evaluated for photocatalytic H2 pro-
duction. The wrapped carbon layer acted as an electron transport bridge between
HTM and W2C, for the reduction of a proton [38]. The wrapped carbon layer boosted
the electron transport from HTM to W2C and reduced the hole transport from W2C
to HTM; this electron hole transfer effectively hindered rapid charge recombination,
represented in (Fig.  8.11). Remarkably, decoration of W2C@C on the HTM
enhanced 21 times high photocatalytic H2 production rate through effective trans-
port of electron and hindering hole transport [24].
Police et  al. demonstrated the formation of wrinkled graphene-wrapped TiO2
nanotubes by simple one pot hydrothermal process. The wrapping of rGO on TiO2
nanotubes resulted in the formation of Ti–O–C bond between rGO and TiO2. The
formation of chemical bond Ti–O–C between rGO and TiO2 resulted in the higher
wavelength absorbance shift. Further, this Ti–O–C bond helped efficient separation
of the photogenerated electron hole from TiO2 to rGO [39, 40]. Above-improved

Fig. 8.11 (a) Schematic


illustration of the charge
transfer and separation in
W2C@C/HTMs hybrid
under simulated solar light
irradiation. (b) Schematic
electronic band structure
and photocatalytic
mechanism for water
reduction. [Copyright 24]
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 229

Fig. 8.12  Effect of GO loading on the photocatalytic activity of rGO–TiO2 composite (a),
Comparison of photocatalytic H2 production of GO, P25 TiO2, bare TiO2 and 5rGO–TiO2 (b), Time
on stream H2 production (c), recycling studies over 5rGO–TiO2 (d). [Copyright 25]

light absorbance and efficient charge carrier separation through Ti-O-C bond formed
between rGO and TiO2 lead to ~13 times higher photocatalytic H2 production and
high stability of photocatalyst even after 5 cycles as shown in (Fig. 8.12) [25].

8.4  Future Direction

The interesting properties of semi-metallic graphene, such as high surface area,


2D-structure, high chemical and thermal stability, high electron mobility, and con-
ductivity, have made it attractive for photocatalytic applications. As discussed in
this chapter, rGO wrapping approach photocatalytic properties of photocatalyst due
to the rapid transport of the electron through rGO layer from the interface of photo-
catalyst and rGO. Besides, it provides solid supports to the photocatalysts and also
functional groups of the rGO acts as adsorption sites for the reactants. It is known
that work function of the graphene is above the hydrogen evolution potential, which
acts as co-catalysts for hydrogen production. Further, the hetero atom doped rGO
230 V. Vinesh et al.

effectively can be utilized in the wrapping approach that provides additional adsorp-


tion sites for hydrogen evolution. Moreover, rGO effectively used as solid-state
mediator in the Z-scheme photocatalysts transport charge carrier transport. These
approaches may lead to the commercial viability of graphene-wrapped photocata-
lyst due to the low processing cost of the rGO.

8.5  Conclusion

In precise, even though numerous approaches have been developed to enhance the
charge carrier separation of semiconductor metal oxide photocatalysts, by introduc-
ing rGO with different metal oxides as support, it significantly enhanced the photo-
generated charge carrier separation due to the chemical attachment of metal oxides
with rGO which resulted in the formation of M–C bonds, which also facilitated the
charge carrier separation apart from its semi-metallic character. Furthermore, the
wrapping approach played a key role in fabricating the nanostructured photocata-
lysts. It was observed that wrapping of graphene oxide to metal oxides not only
enhanced the charge carrier separation but also provided the nucleation sites to
attach the metal oxide semiconductor on graphene surface. Moreover, the rGO-­
wrapped metal oxides enhanced visible light absorption and charge separation effi-
ciency and improved the visible light photocatalytic properties. Wrapping of carbon
material provided intimate uniform heterojunction interface between metal oxides
to help in efficient charge separation and transport through conduction of electrons
and impede the transport of hole to improve photocatalytic activity. Thus, the for-
mation of M–O–C bond in the wrapped rGO-metal oxide benefits in the reduction
of band gap and hinders charge carrier separation.

References

1. Liu, H., Dong, X., Wang, X., Sun, C., Li, J., & Zhu, Z. (2013). A green and direct synthesis of
graphene oxide encapsulated TiO2 core/shell structures with enhanced photoactivity. Chemical
Engineering Journal, 230, 279–285.
2. Tan, L.  L., Chai, S.  P., & Mohamed, A.  R. (2012). Synthesis and applications of graphene
based TiO2 photocatalysts. ChemSusChem. https://doi.org/10.1002/cssc.201200480.
3. Kim, H. I., Moon, G. H., Satoca, D. M., Park, Y., & Choi, W. Y. (2012). Solar photoconversion
using graphene/TiO2 composites: nanographene shell on TiO2 core versus TiO2 nanoparticles
on graphene sheet. Journal of Physical Chemistry C, 116, 1535–1543.
4. Jiang, G.  D., Lin, Z.  F., Chen, C., Zhu, L.  H., Chang, Q., Wang, N., Wei, W., & Tang,
H. Q. (2011). TiO2 nanoparticles assembled on graphene oxide nanosheets with high photo-
catalytic activity for removal of pollutants. Carbon, 49, 2693–2701.
5. Zhang, X. Y., Sun, Y. J., Cui, X. L., & Jiang, Z. Y. (2012). Economics of liquid hydrogen from
water electrolysis. International Journal of Hydrogen Energy, 37, 811–815.
8  Nanostructuring of Hybrid Materials Using Wrapping Approach to Enhance… 231

6. Liu, J.  C., Liu, L., Bai, H.  W., Wang, Y.  J., & Sun, D.  D. (2011). Gram-scale production
of graphene oxide–TiO2 nanorod composites: Towards high-activity photocatalytic materials.
Applied Catalysis B: Environmental, 106, 76–82.
7. Wang, F., & Zhang, K. (2011). Reduced graphene oxide–TiO2 nanocomposite with high pho-
tocatalystic activity for the degradation of rhodamine B. Journal of Molecular Catalysis A:
Chemical, 345, 101–107.
8. Chen, C., Cai, W. M., Long, M. C., Zhou, B. X., Wu, Y. H., Wu, D. Y., & Feng, Y. J. (2010).
Synthesis of visible-light responsive graphene oxide/TiO2 composites with p/n heterojunction.
ACS Nano, 4, 6425–6432.
9. Ni, Y., Wang, W., Huang, W., Lu, C., & Xu, Z. (2014). Graphene strongly wrapped TiO2 for
high-reactive photocatalyst: A new sight for significant application of graphene. Journal of
Colloid and Interface Science, 428, 162–169.
10. Wang, W., Lu, C., Ni, Y., Su, M., & Xu, Z. (2012). Applied Catalysis B: Environmental,
127, 28.
11. Gu, L., Wang, J., Cheng, H., Du, Y., & Han, X. (2012). Chemical Communications, 48, 6978.
12. Vinesh, V., Shaheer, A. R. M., & Neppolian, B. (2019). Reduced graphene oxide (rGO) sup-
ported electron deficient B-doped TiO2 (Au/B-TiO2/rGO) nanocomposite: An efficient visible
light sonophotocatalyst for the degradation of Tetracycline (TC). Ultrasonics Sonochemistry,
50, 302–310.
13. Lavanya, T., Satheesh, K., Dutta, M., Jaya, N. V., & Fukata, N. (2014). Superior photocatalytic
performance of reduced graphene oxide wrapped electrospun anatase mesoporous TiO2 nano-
fibers. Journal of Alloys and Compounds, 615, 643–650.
14. Ganesh Babu, S., Vinoth, R., Surya Narayana, P., Bahnemann, D., & Neppolian, B. (2015).
Reduced graphene oxide wrapped Cu2O supported on C3N4: An efficient visible light respon-
sive semiconductor photocatalyst. APL Materials, 3(10), 104415.
15. Kim, H. I., Moon, G. H., Monllor-Satoca, D., Park, Y., & Choi, W. (2012). Journal of Physical
Chemistry C, 116, 1535–1543.
16. Akhavan, O., & Ghaderi, E. (2013). Nanoscale, 5, 10316–10326.
17. Akhavan, O., Azimirad, R., Safa, S., & Larijani, M. M. (2010). Journal of Materials Chemistry,
20, 7386–7392.
18. Bell, N. J., Ng, Y. H., Du, A., Coster, H., Smith, S. C., & Amal, R. (2011). Journal of Physical
Chemistry C, 115, 6004–6009.
19. Zhai, J., Yu, H., Li, H., Sun, L., Zhang, K., & Yang, H. (2015). Visible-light photocatalytic
activity of graphene oxide-wrapped Bi2WO6 hierarchical microspheres. Applied Surface
Science. https://doi.org/10.1016/j.apsusc.2015.03.100.
20. Stoltzfus, M. W., Woodward, P. M., Seshadri, R., Klepei, J.-H., & Bursten, B. (2007). Inorganic
Chemistry, 46, 3839–3850.
21. Zhang, L. S., Wang, W. Z., Yang, J. O., Chen, Z. G., Zhang, W. Q., & Zhou, L. (2006). Applied
Catalysis A: General, 308, 105–110.
22. Li, H. Q., Cui, Y. M., & Hong, W. S. (2013). Applied Surface Science, 264, 581–588.
23. Yu, Y., Liu, Y., Wu, X.  Q., Weng, Z.  H., Hou, Y., & Wu, L.  S. (2015). Separation and
Purification Technology, 142, 1–7.
24. Yue, X.-Z., et al. (2019). Steering charge kinetics in W2C@ C/TiO2 heterojunction architec-
ture: Efficient solarlight-driven hydrogen generation. Applied Catalysis B: Environmental,
255, 117760.
25. Police, K. R., et al. (2018). Single-step hydrothermal synthesis of wrinkled graphene wrapped
TiO2 nanotubes for photocatalytic hydrogen production and supercapacitor applications.
Materials Research Bulletin, 98, 314–321.
26. Xu, et al. (2019). Interfacial bonded CuCo2O4/TiO2 nanosheets heterostructures for boosting
photocatalytic H2 production. Catalysis Science & Technology.
27. Qin, H., Zhao, X., Zhao, H., Yan, L., & Fan, W. (2019). Well-organized CN-M/CN-U/Pt-TiO2
ternary heterojunction design for boosting photocatalytic H2 production via electronic continu-
ous and directional transmission. Applied Catalysis A: General, 576, 74–84.
232 V. Vinesh et al.

28. Karthik, P., et al. (2017). A visible-light active catechol–metal oxide carbonaceous polymeric
material for enhanced photocatalytic activity. Journal of Materials Chemistry A, 5, 384–396.
29. Ning, et  al. (2016). TiO2/graphene/NiFe-layered double hydroxide nanorod array photo-
anodes for efficient photoelectrochemical water splitting. Energy & Environmental Science, 9,
2633–2643.
30. Park, H.-A., et al. (2017). Nano-photoelectrochemical cell arrays with spatially isolated oxida-
tion and reduction channels. ACS Nano, 11, 2150–2159.
31. Zhang, et  al. (2018). Hierarchical honeycomb Br-, N-codoped TiO2 with enhanced visible-­
light photocatalytic H2 production. ACS Applied Materials & Interfaces, 10, 18796–18804.
32. Hafeez, Y., Lakhera, S.  K., Ashokkumar, M., & Neppolian, B. (2019). Ultrasound assisted
synthesis of reduced graphene oxide (rGO) supported InVO4–TiO2 nanocomposite for efficient
hydrogen production. Ultrasonics Sonochemistry, 53, 1–10.
33. Dou, et al. (2019). Core-shell g-C3N4/Pt/TiO2 nanowires for simultaneous photocatalytic H2
evolution and RhB degradation under visible light irradiation. Catalysis Science & Technology.
34. Hejazi, S., Altomare, M., Mohajernia, S., & Schmuki, P. (2019). Composition-gradients in
sputtered Ti–Au alloys: Site-selective Au-decoration of anodic TiO2 nanotubes for photocata-
lytic H2 evolution. ACS Applied Nano Materials.
35. Xu, et al. (2018). Noble metal-free rGO/TiO2 composite nanofiber with enhanced photocata-
lytic H2-production performance. Applied Surface Science, 434, 620–625.
36. Umer, et al. (2019). Montmorillonite dispersed single wall carbon nanotubes (SWCNTs)/TiO2
heterojunction composite for enhanced dynamic photocatalytic H2 production under visible
light. Applied Clay Science, 174, 110–119.
37. Li, Y., et  al. (2018). Enhanced photocatalytic H2-production activity of C-dots modified
g-C3N4/TiO2 nanosheets composites. Journal of Colloid and Interface Science, 513, 866–876.
38. Tachibana, Y., Vayssieres, L., & Durrant, J. R. (2012). Artificial photosynthesis for solar water-­
splitting. Nature Photonics, 6, 511.
39. Chen, X., Shen, S., Guo, L., & Mao, S. S. (2010). Semiconductor-based photocatalytic hydro-
gen generation. Chemical Reviews, 110, 6503–6570.
40. Babu, S. G., et al. (2015). Influence of electron storing, transferring and shuttling assets of
reduced graphene oxide at the interfacial copper doped TiO2 p–n heterojunction for increased
hydrogen production. Nanoscale, 7, 7849–7857.
Chapter 9
Metal–Organic Frameworks (MOFs)
with Hierarchical Structures for Visible
Light Photocatalysis

P. Karthik and B. Neppolian

9.1  Introduction

Sustainable solar energy storage and conversion can be a promising substitute for
conventional fossil fuels as running out of fossil fuels can lead to a future energy
crisis [1–3]. Despite the high abundance of solar energy, still efficient storage and
conversion are challenging tasks due to the limited efficiency of the available meth-
ods. Many ways have been proposed for effective utilization of solar energy, par-
ticularly in semiconductor photocatalysis for the production of green fuel hydrogen
(H2), conversion of CO2 to useful chemicals, and environmental remediation appli-
cations, including heavy metals reduction and organic pollutants degradation, etc.,
[4–7]. So far, various semiconductors such as metal oxides, metal sulfides, metal
oxychlorides, and organic polymers, etc., have been used as photocatalysts for many
application studies [8–11]. Though the various semiconductor photocatalysts are in
hand, the well-known photocatalyst TiO2 takes precedence over  other photocata-
lysts. This is due to its advantages, including superior stability under harsh experi-
mental conditions, suitable band edge potential, etc. However, the wide band gap
(3.2 eV) is a major concern with TiO2. So far, many methods have been used for the
synthesis of visible light active TiO2, namely, doping of noble/non-noble transition
metals, creating composites with other metal oxides, surface sensitization, etc. [12–
15]. Howbeit, excluding the photocatalysts mentioned above, the metal–organic
frameworks (MOFs) have also gained much attention in visible light semiconductor
photocatalysis.
MOFs are classified under the category of porous materials, formed by the peri-
odical arrangement of a metal cluster and organic linkers through the coordination
interactions [16, 17]. The interesting properties, such as high surface area, tunable

P. Karthik · B. Neppolian (*)


SRM Research Institute, SRM Institute of Science and Technology, Kattankulathur,
Chennai, Tamil Nadu, India
e-mail: neppolib@srmist.edu.in

© Springer Nature Switzerland AG 2021 233


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_9
234 P. Karthik and B. Neppolian

architecture, high porosity, etc., widen their uses in diverse research  fields.
Especially, MOFs are a potential candidate for gas storage, sensor, and catalysis
applications. However, recently their contribution in photocatalysis is similar to
inorganic semiconductors due to the following advantages: (i) the high porosity of
MOFs facilitates the adsorption of reacting molecules on the surface, which
increases the rate of reaction.; (ii) the tunable structure of MOFs provides a chance
to extend their light absorption in a wide range; (iii) the high crystallinity and less
structural defects reduce the rate of electron–hole pair recombination; and (iv) the
flexibility allows the introduction of co-catalysts into pores of MOFs, which could
enhance the spatial separation of electron–hole pairs [18]. Inspired by these merits,
extensive efforts have been adopted to develop MOF-based materials for various
photocatalytic applications.
Besides, the hierarchical nanostructures of materials play a vital role in photoca-
talysis. The rate of photogenerated electron–hole pairs can be improved by the hier-
archical nanostructure of the materials, which can enhance the photocatalytic
activity [19]. Likewise, the fabrication of MOFs with different morphology and
pore size can exhibit considerable enhancement in photocatalytic activities. In this
present chapter, we will discuss the roles and effects of hierarchical nanostructures
of MOFs in photocatalysis.

9.2  Rational Fabrication of MOFs Photocatalyst

Though the MOFs emerge as promising materials for many applications since 1990,
it has been considered to be a potential material for photocatalytic applications after
the year of 2000. Recently, MOFs are widely used for photocatalytic applications as
they have semiconducting behavior that is analogous with conventional inorganic
metal oxide semiconductors. The MOFs are formed due to the bridging of inorganic
metal nodes and organic ligands or linkers (Fig. 9.1) [20]. Due to the broad absorp-
tion in UV and visible region with band edge that lies in the typical semiconductor
like band gap values, MOFs are classified as semiconductors. But it is not an appro-
priate point to classify the MOFs as the semiconductor [21], because the organic
linkers of MOFs can absorb light both UV and visible region due to π–π* or ligand
to metal charge transfer process (LMCT) as similar to that of metal complex. Some
of the MOFs hold insulating behavior due to the weak overlap of frontier orbitals
that make the electronic states localized, not delocalized [21]. The optical response
of MOFs can be tuned by the building blocks such as organic linkers. With the
knowledge of the chromophoric nature of organic building blocks, the superior
light-harvesting MOFs can be constructed. By choosing highly conjugated organic
linkers to fabricate MOFs, the optical response can be enhanced, offering the vast
platform for fabrication of MOFs as a photocatalyst [22]. In addition, the multi-­
components of MOFs facilitate the synergetic action in photocatalytic reactions as
9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 235

Fig. 9.1  Fabrication of MOFs {Reproduced with permission {Reproduced with permission from
ref. [17], Copyright (2019), American Chemical Society} from ref. [5], Copyright (2017), Elsevier}

the proximity between components leads to the photogenerated carrier separation.


Moreover, though the optical absorbance of MOFs can be enhanced by organic
building blocks, the charge carrier recombination is another complex problem in
MOFs photocatalysts [23]. Various co-catalysts have been integrated with MOFs to
enhance the charge carrier rate in MOFs photocatalysts. Particularly, introduction of
Pt and Au as co-catalysts in MOFs significantly improves charge carrier separation
rate. For instance, a report by Jiang and co-workers [24] stated that the location of
Pt nanoparticles on UiO-66-NH2, considerably affects the electron–hole pair recom-
bination. They have synthesized the two different Pt-UiO-66-NH2 MOFs and depos-
ited the Pt nanoparticles in different locations by solvothermal and in situ growth
methods. It was noted that Pt-loaded MOF (Pt@ Pt-UiO-66-NH2) synthesized by in
situ growth method showed 5 times enhanced photocatalytic H2 production activity
than the Pt-MOF ((Pt/Pt-UiO-66-NH2) synthesized by solvothermal method. The
location of Pt nanoparticles was mainly attributed to the superior activity of Pt@
Pt-UiO-66-NH2. During the in situ synthesis of Pt@Pt-UiO-66-NH2 MOF, the Pt
nanoparticles are inserted into the pores of MOF; as a result, the proximity between
MOF and Pt nanoparticles was increased [24]. Moreover, the molecular complex
can also be integrated with MOFs as a co-catalyst. Nasalevich et al. introduced the
cobaloxime as a molecular co-catalyst with NH2-MIL-125(Ti) MOF, which exhib-
ited the enhanced H2 production activity. In order to synthesize cobaloxime-NH2-­
MIL-125(Ti) MOF composite, firstly, they have introduced flexible LH2(DOH)2pn
organic ligands into pores of NH2-MIL-125(Ti) MOF. Next, CoBr2 was added to it
under aerobic condition, which formed the complex with formerly added
LH2(DOH)2pn organic ligands [25].
236 P. Karthik and B. Neppolian

9.3  Various Methods for MOFs Synthesis

As shown in Fig.  9.2, several methods have been employed to synthesize MOFs
with different experimental conditions. The simple method to synthesize MOFs is
the mixing of organic building blocks and metals in a liquid medium. As a result of
mixing in a liquid medium, the MOF powder will be formed [26]. Further, the prod-
uct can be separated by filtration and dried to get solvent-free MOF powder.
However, the most popular methods for the synthesis of MOFs are hydrothermal
and solvothermal. For these methods, the solvents with decent solubility such as
acetone, DMF, DMSO, ethanol, methanol, acetonitrile, etc. are used to dissolve the
organic linkers. Notably, the mixture of solvents can be used if the initial reactants
possess different solubility. After mixing all the starting materials, the hydrothermal
or solvothermal reaction can be carried out in a Teflon-lined autoclave or glass vial.
The selection of Teflon-lined autoclave and glass vial is based on the reaction tem-
perature. If the low temperature can induce the MOFs formation, glass vial can be
used for the hydrothermal/solvothermal reaction. Besides, if the MOFs formation
reaction takes place at high temperature, Teflon-lined autoclaves can be used for the
MOFs synthesis [27]. On the other hand, the hydrothermal method, i.e., the use of
water as a solvent for MOFs synthesis is considered to be an  environmentally
friendly approach, which may increase the interest of MOFs synthesis without using
any organic solvents. Some of the literature reported the synthesis of MOFs in water
medium. For example, the carboxylate organic salts were used for the synthesis of
carboxylate-based MOFs. Hence, the solubility of such kind of organic linkers is
high in water medium that enables the MOFs formation [28]. However, the time
span of the hydro−/solvothermal reactions to form high crystalline MOFs is long.

Fig. 9.2  Various synthesis methods of MOFs


9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 237

So, other alternative methods, namely, microwave-assisted, electrochemical, and


mechanochemical methods, were also proposed for the synthesis of MOFs in a short
span of time with high crystallinity. With these methods, it is possible to synthesize
high crystalline MOFs with a short span of time. In addition, the following charac-
teristics of solvents such as solubility, redox potential, and reactivity should be
minded while choosing a solvent for the MOF synthesis.

9.4  MOFs for Visible Light Photocatalysis

The continuous aim is to develop visible light active photocatalytic materials that
have been utilized to harvest 45% of visible photons in solar light [29]. Therefore,
it is foremost essential to design and develop visible photons active photocatalysts.
In this regard, many methods have been devoted to making inorganic semiconduc-
tors as visible light photocatalysts. Excluding the conventional inorganic semicon-
ductor photocatalysts, MOFs are growing as a new class of hybrid photocatalysts.
Some of the properties of MOFs are similar to zeolites and metal oxides, which
make them a potential candidate for photocatalytic applications. Due to the interest-
ing structural properties of MOFs, which has been sought for the development of
light-responsive materials in photocatalysis. In this section, we discuss the visible
light photocatalytic activity of MOFs with hierarchical structures.
In general, the rate of electron-hole pair separation plays a crucial role in photo-
catalysis. Studying the fundamental insights on charge carrier migration and separa-
tion in photocatalysis could direct the rational design of MOFs photocatalysts [21].
Various methods have reported boosting electron–hole separation in MOF photoca-
talysis. Besides, the fabrication of MOF-based composite by introducing co-­
catalysts has considered an effective way of enhancing the electron–hole separation
rate. As similar to conventional metal oxide photocatalyst, metal nanoparticles can
be introduced with MOFs, where metal nanoparticles behave as electron acceptors.
However, Karthik et  al. fabricated the reduced graphene oxide (rGO) supported
noble metal-free Ti-MOF with a high electron–hole pair rate for visible light photo-
catalytic H2 production (Fig. 9.3) [29]. Notably, the cubic structure of Ti-MOF did
not collapse by rGO introduction. After the post-synthetic rGO introduction, the
cubic structure of Ti-MOF was retained, which reveals that there is no influence of
rGO introduction on Ti-MOF.  However, in contrast to the structure, a significant
favorable influence was observed in the electron–hole separation rate. This is an
interesting enhancement in charge carrier separation, which was attributed by π–π
interactions between rGO and Ti-MOF. The direct overlapping of pi orbitals of both
rGO and Ti-MOF facilitated the rapid migration of photogenerated charge carriers.
Apart from this, ninefold enhanced photocatalytic H2 production rate was obtained
with rGO/Ti-MOF than bare Ti-MOF with excellent recyclability (Fig. 9.4). This
significantly enhanced photocatalytic activity was mainly due to rapid charge car-
rier separation through π–π interaction and improved visible light absorption of
rGO/Ti-MOF photocatalyst. Thus, it isclear from this work that the structure of
238 P. Karthik and B. Neppolian

Fig. 9.3  TEM images of (a)  NH2-MIL-125 (Ti) and (b) rGO/NH2-MIL-125 (Ti) MOFs. (c–h)
HAADF-STEM imges of rGO/NH2-MIL-125 (Ti) MOF {Reproduced with permission from ref.
[29], Copyright (2019), American Chemical Society}

Fig. 9.4  Photocatalytic H2 production activity of (a) NH2-MIL-125 (Ti) and rGO/ H2-MIL-125
(Ti) MOFs and (b) recycle test of rGO/ MIL-125 (Ti) MOF{Reproduced with permission from ref.
[29], Copyright (2019), American Chemical Society}
9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 239

Ti-MOF is quite stable against post-synthetic introduction of rGO and the π–π inter-
action facilitates the rapid electron–hole pair separation.
Zhang et al. prepared the Ru-MOF with a nanoflower structure for visible light-­
driven photocatalytic CO2 reduction. The nanoflower structure of Ru-MOF exhib-
ited good stability against photocatalytic CO2 reduction reaction [30]. Moreover,
this work brings out a relationship between the structure of MOF and photocatalytic
activity. They have noted that nanoflower Ru-MOF demonstrated a BET surface
area of 8.08 m2g−1 which was higher than the micro-crystal MOF (1.33 m2g−1). As a
result of enhanced surface area, the catalytic active sites for CO2 reduction were also
increased considerably. Compared to solid micro-crystal MOF, nanoflower Ru-MOF
showed a 150% enhanced photocatalytic CO2 reduction under the same experimen-
tal conditions. This considerably enhanced photocatalytic activity was ascribed to
the nanoflower structure of Ru-MOF (Fig.  9.5). Further, the excited state energy
transfer efficiency was also high for Ru-MOF with a nanoflower structure. All these
factors have contributed to the efficient photocatalytic CO2 reduction under visi-
ble light.
In addition to the flower-like morphology of MOF, 2D structure of MOFs could
also play a vital role in photocatalytic activity. For example, Peng and his co-­
workers reported the Ni-based 2D conductive MOF as an efficient co-catalyst for
visible light-driven CO2 reduction [31]. The 2D structure of Ni-MOF possessed
high conductivity due to the honeycomb structure and planer Ni-N4 sites. The con-
ductivity of polycrystalline films and pressed pellets was found to be 40 and 5000
Sm−1, respectively. On the other hand, they have used this Ni-MOF as a co-catalyst
along with Ru(bpy)3]2+photosensitizer to separate the photogenerated charge carri-
ers. A high CO2 conversion rate of 3.45 × 104 μmol·g−1 h−1 was achieved with this
catalytic system. Moreover, the selectivity of CO2 conversion was also controlled by
this catalyst. In this work, they have obtained 97% selectivity for CO2 reduction
reaction under visible light. The improved CO2 conversion rate and high selectivity

Fig. 9.5  Photocatalytic CO2 conversion activity of Ru-MOF different morphologies {Reproduced
with permission from ref. [30], Copyright (2019), Royal Society of Chemistry}
240 P. Karthik and B. Neppolian

facilitated by 2D structure and planer Ni–N4 sites of the catalyst [31]. Also, MOFs
can be a supporting material to incorporate metal nanoparticles. The incorporation
of metal nanoparticles with MOFs could enhance photocatalytic activity.
Re-containing UiO-67 MOF with octahedral morphology was synthesized by Choi
et al. for visible light CO2 reduction. On this MOF, Ag nanocubes were introduced
to enhance CO2 conversion efficiency [29]. The Ag nanocubes introduced Ru-MOF
demonstrated sevenfold enhanced photocatalytic CO2 to CO conversion efficiency
under visible light illumination. The plasmonic Ag nanocubes and photo-responsive
Ru centers synergistically activated the CO2 conversion [32]. Sun and his group
prepared the ZIF-67 MOFs with different morphologies by a solvent-induced
method [33]. They have studied the effect of morphologies of ZIF-67 MOFs on vis-
ible light photocatalytic CO2 reduction. For this purpose, three different morpholo-
gies of MOFs such as rhombic dodecahedral (ZIF-67-1), pitaya-like morphology
(ZIF-67-2), and two dimensional leaf-like morphology (ZIF-67-3) have prepared by
varying experimental conditions. The visible light photocatalytic activities of all the
prepared ZIF-67 MOFs with different morphologies were studied along with
[Ru(bpy)3]2+ as the main catalyst. From the experimental results, they declared that
ZIF-67 with a leaf-like morphology showed the highest visible light photocatalytic
CO2 conversion activity than other morphologies. The high activity of ZIF-67 with
a leaf-like morphology was due to selective adsorption of CO2 on catalyst surface
[33]. This work gives a clear idea about effect of morphology on photocatalytic
activity of MOFs. Overall, the various roles of MOFs and their morphology-­
dependent photocatalytic activity were discussed in this section.
Aside from that, many MOF-based catalysts were reported with unique crystal
structure for visible light photocatalysis [34]. For instance, Cd(II)-based MOF with
an infinite 3D network was synthesized for multifunctional applications, including
visible light photocatalytic MO degradation and Cr(VI) reduction [35]. The photo-
catalytic activities of this MOF were attributed to broad visible light absorption.
Also, this MOF possesses high dye adsorption capability due to the presence of
peripheral hydroxyl groups. On the other hand, zinc-based MOF with 2D structure
was synthesized by Kaur et al. for visible light-driven photocatalytic Cr(VI) reduc-
tion reaction and recognition [36].

9.5  V
 isible Light Photocatalysis of MOF-Derived
Hierarchical Structured Metal Oxides

In addition to the direct uses of MOF in visible light photocatalysis, it has been used
as a sacrificial template to synthesis hierarchical structured metal oxides for various
applications, in particular, energy and environmental remediation applications. It is
well known that semiconductor metal oxides are important catalytic materials in
photocatalysis. The key advantage of MOFs has been used as a sacrificial template
for the preparation of porous structured metal oxides. The metal oxides with porous
9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 241

structures possess reasonable surface areas that provide more active sites for photo-
catalytic activity. For instance, Lin et al. proposed a facile two-step method to syn-
thesis Fe2O3@TiO2 nanocomposite in which they have used MIL-101(Fe) as a
sacrificial template [37]. Initially, MIL-101(Fe) MOF was coated with TiO2
nanoparticle by acid hydrolysis method. Subsequently, the resultant amorphous
TiO2 /MIL-101(Fe) MOF composite was subjected for calcination under air atmo-
sphere to obtain Fe2O3@TiO2 nanostructure. The photocatalytic activity of prepared
Fe2O3@TiO2 nanostructure was studied towards photocatalytic H2 production under
visible light illumination. The Fe2O3@TiO2 structure exhibited an enhanced rate of
H2 production than its counterparts. In similar way, Xiong’s research group pre-
pared the Cu/TiO2 composite with a hollow structure that was successfully prepared
from Cu3(BTC)2 MOF octahedral microcrystals template for photocatalytic H2 pro-
duction via water splitting. The better octahedral-shaped shells were obtained by
this method compared to other calcination methods [38]. This report concluded that
the MOF-based templated synthesis offers metal oxides with the desired morphol-
ogy. Consequently, the metal oxide heterostructures can be synthesized by MOF-­
based templated synthesis. For example, Pham et  al. prepared hollow
Fe2O3–TiO2–PtOx heterojunction photocatalyst by nanosized MIL-88B MOF as a
template. The prepared Fe2O3-TiO2–PtOx heterojunction demonstrated improved
photocatalytic H2 production under visible light illumination [39]. All the above
studies give significances of MOFs in the templated synthesis of metal oxides for
visible light photocatalysis. Some of the metal oxide synthesis by MOF-based tem-
plated methods and their visible light photocatalytic applications are summarized in
Table 9.1.

Table 9.1  Visible light photocatalytic activities of MOF-derived nanomaterials


Sl MOF Light Photocatalytic H2
No Metal oxide photocatalyst template source production activity Ref.
1 Co-Zn0.5Cd0.5S ZnCo-ZIF Visible 17,360 μmol/h/g [40]
light
2 NiS/Zn0.5Cd0.5S Ni/ Visible 16,780 μmol/h/g [41]
ZnCd-MOF light
4 CdS/ZCO ZnCo-ZIF Visible 3978.6 μmol/h/g [42]
light
5 Yolk–shell CdS Cd–Fe–PBA Visible 3051.4 μmol/h/g [43]
light
6 Hollow MIL-88B@ Visible 1100 μmol/h/g [39]
Fe2O3– TiO2–PtOx TiO2 light
7 ZnO/Au Au/ZIF-8 Visible 29.8 μmol/h/g [44]
light
8 Fe2O3@TiO2 MIL-101@ Visible 625 μmol/h/g [37]
TiO2 light
9 FeOx-carbonaceous MIL-88B/ Visible 264.1 μmol/h/g [45]
composites rGO light
242 P. Karthik and B. Neppolian

9.6  V
 isible Light Photocatalysis of MOF-Derived
Hierarchical Structured Metal Sulfides

Metal sulfides are also widely used as photocatalysts for many applications. In spe-
cific, cadmium sulfide (CdS) is one of the promising photocatalysts for solar energy
conversion applications due to its visible light-driven bandgap (2.4 eV). However,
the rapid recombination of photogenerated charge carriers and high photocorrosion
tendency strongly affects their practical uses. Among these issues, the rapid charge
carrier recombination can be minimized by introducing co-catalysts and increasing
surface area of CdS that can improve the photocatalytic activity. Various synthesis
methods have been reported to enhance surface area of CdS. However, MOF-derived
synthesis of CdS is considered as a facile route to enhance the surface area. For
example, Jiang and Xiao prepared porous CdS with a hierarchical structure by
MOF-templated method. They have prepared cadmium oxide (CdO) from the MOF
template and the subsequent nanocasting result in the formation of porous
CdS. Notably, CdS with a surface area of 119 m2 g−1 was obtained by this method.
The photocatalytic efficiency of porous CdS was tested through photocatalytic H2
production reaction. The porous CdS showed the improved rate of H2 production
than bulk CdS under visible light illumination. This enhanced activity was facili-
tated by high surface area and porosity of CdS [46]. Furthermore, Wang et al. syn-
thesized yolk-shell-structured CdS nanomaterial via a tow step MOF template-based
method for visible light photocatalytic H2 production. The Cd–Fe Prussian blue
analogues (Cd–Fe–PBA) was used as a template for yolk-shell-structured CdS syn-
thesis. The synthesized CdS nanomaterial exhibited high surface area, which
enhanced the photocatalytic H2 production activity [43]. On the other hand,
Co-based sulfide also was synthesized by MOF-templated method. Kim and his
research group synthesized slice-type hollow-structured Co4S3 from Co-MOF. The
composite of Co4S3 and CdS demonstrated improved photocatalytic H2 production
in the presence of simulated solar light. The hollow structure of prepared nanocom-
posite induced the effective charge carrier separation [47]. Thus, the abovemen-
tioned report clearly proposes the uses of MOFs in the synthesis of metal sulfides
with high porous and surface area for visible light photocatalysis.

9.7  Conclusion

In this chapter, we have discussed the preparation methods, properties, and visible
light photocatalytic activity of MOFs with hierarchical structures. The properties,
such as surface area and pores size of MOFs, were strongly influenced by different
structures, which offered improved visible-light photocatalytic performances.
Moreover, the hierarchical structured MOFs were not only used as the main catalyst
for photocatalytic reactions but also used co-catalysts. The dimension of MOFs
played a major role in conductivity. MOF with a 2D structure exhibited a high
9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 243

conductivity and showed enhanced photocatalytic activity. Also, we have discussed


MOF-derived metal oxides and sulfides nanomaterials and their visible-light photo-
catalytic activity. Synthesis of metal oxide and sulfides photocatalysts by MOF-­
derived approach provides a high surface area, which is beneficial for photocatalytic
activity. In comparison to bulk materials, MOF-derived metal oxide and sulfide
demonstrated superior photocatalytic activity. Thus, this chapter combines the role
of hierarchical structures of MOFs on visible light photocatalysis and uses of MOFs
for the fabrication of nanomaterials, including metal oxides and sulfides
photocatalysts.

References

1. Grasemann, M., & Laurenczy, G. (2012). Formic acid as a hydrogen source–recent develop-
ments and future trends. Energy & Environmental Science, 5, 8171–8181.
2. Wang, X., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J.  M., Domen, K., &
Antonietti, M. (2009). A metal-free polymeric photocatalyst for hydrogen production from
water under visible light. Nature Materials, 8, 76.
3. Alarawi, A., Ramalingam, V., & He, J.-H. (2019). Recent advances in emerging single atom
confined two-dimensional materials for water splitting applications. Materials Today Energy,
11, 1–23.
4. Karthik, P., Vinoth, R., Selvam, P., Balaraman, E., Navaneethan, M., Hayakawa, Y., &
Neppolian, B. (2017). A visible-light active catechol–metal oxide carbonaceous polymeric
material for enhanced photocatalytic activity. Journal of Materials Chemistry A, 5, 384–396.
5. Karthik, P., Balaraman, E., & Neppolian, B. (2018). Efficient solar light-driven H2 produc-
tion: post-synthetic encapsulation of a Cu2O co-catalyst in a metal–organic framework (MOF)
for boosting the effective charge carrier separation. Catalysis Science & Technology, 8,
3286–3294.
6. Babu, S. G., Karthik, P., John, M. C., Lakhera, S. K., Ashokkumar, M., Khim, J., & Neppolian,
B. (2019). Synergistic effect of sono-photocatalytic process for the degradation of organic pol-
lutants using CuO-TiO2/rGO. Ultrasonics Sonochemistry, 50, 218–223.
7. Ding, M., & Jiang, H.-L. (2018). Incorporation of imidazolium-based poly (ionic liquid) s into
a metal–organic framework for CO2 capture and conversion. ACS Catalysis, 8, 3194–3201.
8. Vinoth, R., Karthik, P., Devan, K., Neppolian, B., & Ashokkumar, M. (2017). TiO2–NiO p–n
nanocomposite with enhanced sonophotocatalytic activity under diffused sunlight. Ultrasonics
Sonochemistry, 35, 655–663.
9. Vinoth, R., Babu, S. G., Ramachandran, R., & Neppolian, B. (2017). Bismuth oxyiodide incor-
porated reduced graphene oxide nanocomposite material as an efficient photocatalyst for vis-
ible light assisted degradation of organic pollutants. Applied Surface Science, 418, 163–170.
10. Li, Z., Meng, X., & Zhang, Z. (2018). Recent development on MoS2-based photocatalysis: A
review. Journal of Photochemistry and Photobiology, 35, 39–55.
11. Karthik, P., Kumar, T. N., & Neppolian, B. (2019). Redox couple mediated charge carrier sepa-
ration in g-C3N4/CuO photocatalyst for enhanced photocatalytic H2 production. International
Journal of Hydrogen Energy.
12. Linsebigler, A. L., Lu, G., & Yates, J. T., Jr. (1995). Photocatalysis on TiO2 surfaces: Principles,
mechanisms, and selected results. Chemical Reviews, 95, 735–758.
13. Xu, C., Ravi Anusuyadevi, P., Aymonier, C., Luque, R., & Marre, S. (2019). Nanostructured
materials for photocatalysis. Chemical Society Reviews, 48, 3868–3902.
244 P. Karthik and B. Neppolian

14. Kumar, A., Kumar, K., & Krishnan, V. (2019). Sunlight driven methanol oxidation by aniso-
tropic plasmonic Au nanostructures supported on amorphous titania: Influence of morphology
on photocatalytic activity. Materials Letters, 245, 45–48.
15. Kumar, A., Sharma, V., Kumar, S., Kumar, A., & Krishnan, V. (2018). Towards utilization of
full solar light spectrum using green plasmonic Au–TiOx photocatalyst at ambient conditions.
Surfaces and Interfaces, 11, 98–106.
16. Jiao, L., Wang, Y., Jiang, H. L., & Xu, Q. (2018). Metal–organic frameworks as platforms for
catalytic applications. Journal of Advanced Materials, 30, 1703663.
17. Karthik, P., Pandikumar, A., Preeyanghaa, M., Kowsalya, M., & Neppolian, B. (2017). Amino-­
functionalized MIL-101 (Fe) metal–organic framework as a viable fluorescent probe for
nitroaromatic compounds. Mikrochimica Acta, 184, 2265–2273.
18. Zeng, L., Guo, X., He, C., & Duan, C. (2016). Metal–organic frameworks: Versatile materials
for heterogeneous photocatalysis. ACS Catalysis, 6, 7935–7947.
19. Zhan, W., Sun, L., & Han, X. (2019). Recent progress on engineering highly efficient

porous semiconductor photocatalysts derived from metal–organic frameworks. Nano-Micro
Letters, 11, 1.
20. Cui, P., Wang, P., Zhao, Y., & Sun, W.-Y. (2019). Fabrication of desired metal–organic frame-
works via postsynthetic exchange and sequential linker installation. Crystal Growth & Design,
19, 1454–1470.
21. Dhakshinamoorthy, A., Asiri, A. M., & Garcia, H. (2016). Metal–organic framework (MOF)
compounds: Photocatalysts for redox reactions and solar fuel production. Angewandte Chemie,
55, 5414–5445.
22. Zhang, T., & Lin, W. (2014). Metal–organic frameworks for artificial photosynthesis and pho-
tocatalysis. Chemical Society Reviews, 43, 5982–5993.
23. Xiao, J.-D., & Jiang, H.-L. (2018). Metal–organic frameworks for photocatalysis and photo-
thermal catalysis. Accounts of Chemical Research, 52, 356–366.
24. Xiao, J. D., Shang, Q., Xiong, Y., Zhang, Q., Luo, Y., Yu, S. H., & Jiang, H. L. (2016). Boosting
photocatalytic hydrogen production of a metal–organic framework decorated with platinum
nanoparticles: The platinum location matters. Angewandte Chemie, 55, 9389–9393.
25. Nasalevich, M. A., Becker, R., Ramos-Fernandez, E. V., Castellanos, S., Veber, S. L., Fedin,
M. V., Kapteijn, F., Reek, J. N., Van Der Vlugt, J., & Gascon, J. (2015). Co@NH2-MIL-125
(Ti): Cobaloxime-derived metal–organic framework-based composite for light-driven H2 pro-
duction. Energy & Environmental Science, 8, 364–375.
26. Dey, C., Kundu, T., Biswal, B.  P., Mallick, A., & Banerjee, R. (2014). Crystalline metal–
organic frameworks (MOFs): Synthesis, structure and function. Acta Crystallographica
Section B: Structural Science, Crystal Engineering and Materials, 70, 3–10.
27. Hashemi, B., Zohrabi, P., Raza, N., & Kim, K.-H. (2017). Metal–organic frameworks as
advanced sorbents for the extraction and determination of pollutants from environmental, bio-
logical, and food media. TrAC Trends in Analytical Chemistry, 97, 65–82.
28. Etacheri, V., Di Valentin, C., Schneider, J., Bahnemann, D., & Pillai, S. C. (2015). Visible-­
light activation of TiO2 photocatalysts: Advances in theory and experiments. Journal of
Photochemistry and Photobiology, 25, 1–29.
29. Karthik, P., Vinoth, R., Zhang, P., Choi, W., Balaraman, E., & Neppolian, B. (2018). π–π
Interaction between metal–organic framework and reduced graphene oxide for visible-light
photocatalytic H2 production. ACS Applied Energy Materials, 1, 1913–1923.
30. Zhang, S., Li, L., Zhao, S., Sun, Z., Hong, M., & Luo, J. (2015). Hierarchical metal–organic
framework nanoflowers for effective CO2 transformation driven by visible light. Journal of
Materials Chemistry A, 3, 15764–15768.
31. Zhu, W., Zhang, C., Li, Q., Xiong, L., Chen, R., Wan, X., Wang, Z., Chen, W., Deng, Z., &
Peng, Y. (2018). Selective reduction of CO2 by conductive MOF nanosheets as an efficient co-­
catalyst under visible light illumination. Applied Catalysis B: Environmental, 238, 339–345.
32. Choi, K. M., Kim, D., Rungtaweevoranit, B., Trickett, C. A., Barmanbek, J. T. D., Alshammari,
A.  S., Yang, P., & Yaghi, O.  M. (2016). Plasmon-enhanced photocatalytic CO2 conversion
9  Metal–Organic Frameworks (MOFs) with Hierarchical Structures for Visible Light… 245

within metal–organic frameworks under visible light. Journal of the American Chemical
Society, 139, 356–362.
33. Wang, M., Liu, J., Guo, C., Gao, X., Gong, C., Wang, Y., Liu, B., Li, X., Gurzadyan, G. G., &
Sun, L. (2018). Metal–organic frameworks (ZIF-67) as efficient cocatalysts for photocatalytic
reduction of CO2: The role of the morphology effect. Journal of Materials Chemistry A, 6,
4768–4775.
34. Kaur, H., Kumar, A., Koner, R.  R., & Krishnan, V. (2020). Metal–organic frame-

works for photocatalytic degradation of pollutants. Elsevier. https://doi.org/10.1016/
B978-0-12-818598-8.00006-7.
35. Kaur, H., Kumar, R., Kumar, A., Krishnan, V., & Koner, R. R. (2019). Trifunctional metal–
organic platform for environmental remediation: Structural features with peripheral hydroxyl
groups facilitate adsorption, degradation and reduction processes. Dalton Transactions, 48,
915–927.
36. Kaur, H., Sinha, S., Krishnan, V., & Koner, R. R. (2020). Photocatalytic reduction and recog-
nition of Cr(VI): New Zn(II)-Based metal–organic framework as catalytic surface. Industrial
and Engineering Chemistry Research, 59, 8538–8550.
37. Zhang, Y., Huang, J., & Ding, Y. (2016). Porous Co3O4/CuO hollow polyhedral nanocages
derived from metal–organic frameworks with heterojunctions as efficient photocatalytic water
oxidation catalysts. Applied Catalysis B: Environmental, 198, 447–456.
38. Li, R., Wu, S., Wan, X., Xu, H., & Xiong, Y. (2016). Cu/TiO2 octahedral-shell photocata-
lysts derived from metal–organic framework@ semiconductor hybrid structures. Inorganic
Chemistry Frontiers, 3, 104–110.
39. Pham, M.-H., Dinh, C.-T., Vuong, G.-T., Ta, N.-D., & Do, T.-O. (2014). Visible light induced
hydrogen generation using a hollow photocatalyst with two cocatalysts separated on two sur-
face sides. Physical Chemistry Chemical Physics, 16, 5937–5941.
40. Tang, X., Zhao, J.-H., Li, Y.-H., Zhou, Z.-J., Li, K., Liu, F.-T., & Lan, Y.-Q. (2017). Co-Doped
Zn 1− x Cd x S nanocrystals from metal–organic framework precursors: porous microstructure
and efficient photocatalytic hydrogen evolution. Dalton Transactions, 46, 10553–10557.
41. Zhao, X., Feng, J., Liu, J., Shi, W., Yang, G., Wang, G. C., & Cheng, P. (2018). An efficient,
visible-light-driven, hydrogen evolution catalyst NiS/ZnxCd1− xS nanocrystal derived from a
metal–organic framework. Angewandte Chemie, 130, 9938–9942.
42. Chen, W., Fang, J., Zhang, Y., Chen, G., Zhao, S., Zhang, C., Xu, R., Bao, J., Zhou, Y., & Xiang,
X. (2018). CdS nanosphere-decorated hollow polyhedral ZCO derived from a metal–organic
framework (MOF) for effective photocatalytic water evolution. Nanoscale, 10, 4463–4474.
43. Su, Y., Ao, D., Liu, H., & Wang, Y. (2017). MOF-derived yolk–shell CdS microcubes with
enhanced visible-light photocatalytic activity and stability for hydrogen evolution. Journal of
Materials Chemistry A, 5, 8680–8689.
44. He, L., Li, L., Wang, T., Gao, H., Li, G., Wu, X., Su, Z., & Wang, C. (2014). Fabrication of Au/
ZnO nanoparticles derived from ZIF-8 with visible light photocatalytic hydrogen production
and degradation dye activities. Dalton Transactions, 43, 16981–16985.
45. Yao, J., Chen, J., Shen, K., & Li, Y. (2018). Phase-controllable synthesis of MOF-templated
maghemite–carbonaceous composites for efficient photocatalytic hydrogen production.
Journal of Materials Chemistry A, 6, 3571–3582.
46. Xiao, J. D., & Jiang, H. L. (2017). Thermally stable metal–organic framework-templated syn-
thesis of hierarchically porous metal sulfides: Enhanced photocatalytic hydrogen production.
Small, 13, 1700632.
47. Kumar, D.  P., Park, H., Kim, E.  H., Hong, S., Gopannagari, M., Reddy, D.  A., & Kim,
T.  K. (2018). Noble metal-free metal–organic framework-derived onion slice-type hollow
cobalt sulfide nanostructures: Enhanced activity of CdS for improving photocatalytic hydro-
gen production. Applied Catalysis B: Environmental, 224, 230–238.
Chapter 10
Soil Remediation by Zero-Valent Iron
Nanoparticles for Organic Pollutant
Elimination

Marco Stoller, Luca Di Palma, and Giorgio Vilardi

10.1  Introduction

Nanotechnology studies the manipulation of matter on an atomic, molecular, and


supramolecular scale. The definition generally accepted is that established by the
National Nanotechnology Initiative, which defines nanotechnology as the manipu-
lation of matter with at least one dimension sized from 1 to 100 nm. At this scale,
quantum mechanical effects become important and special properties of matter,
which occur only below the given size threshold.
Nanotechnology as defined only by size results to be very broad. Indeed, nano-
technologies are possible by the interaction of different sciences such as surface
science, organic chemistry, molecular biology, semiconductor physics, energy stor-
age, microfabrication, and molecular engineering. The possible applications are
equally widespread, ranging from extensions of conventional device physics to
completely new approaches based upon molecular self-assembly and from develop-
ing new materials with dimensions on the nanoscale to direct control of matter on
the atomic scale, in the field of medical and human health, pharmacology, cosmet-
ics, food industry, textiles, nutraceuticals, and environmental applications, as
listed below.
Nanomaterials for radioactive waste clean-up in water: The use of titanate nano-
fibers appears to be efficient to absorb radioactive ions from water. Researchers
have also reported that the unique structural properties of titanate nanotubes and
nanofibers make them superior materials for removal of radioactive cesium and
iodine ions in water.
Nanotechnology-based solutions for oil spills: Conventional clean-up techniques
are not adequate to solve the problem of massive oil spills. In recent years,

M. Stoller (*) · L. Di Palma · G. Vilardi


Department of Chemical Engineering Materials Environment, Sapienza University of Rome,
Rome, Italy
e-mail: marco.stoller@uniroma1.it

© Springer Nature Switzerland AG 2021 247


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_10
248 M. Stoller et al.

nanotechnology has emerged as a potential source of novel solutions to many of the


world’s outstanding problems. Although the application of nanotechnology for oil
spill clean-up is still in its nascent stage, it offers great promise for the future. In the
last couple of years, there has been particularly growing interest worldwide in
exploring ways of finding suitable solutions to clean up oil spills through use of
nanomaterials.
Desalination: Nanotechnology-based water purification devices have the poten-
tial to transform the field of desalination, for instance by using the ion concentration
polarization phenomenon.
Treatment of brackish waters: Another relatively new method of purifying brack-
ish water is capacitive deionization (CDI) technology. The advantages of CDI are
that it has no secondary pollution, is cost-effective, and is energy efficient.
Nanotechnology researchers have developed a CDI application that uses graphene-­
like nanoflakes as electrodes for capacitive deionization. They found that the gra-
phene electrodes resulted in a better CDI performance than the conventionally used
activated carbon materials.
Water applications: The potential impact areas for nanotechnology in water
applications are divided into three categories, that is, sensing and detection, pollu-
tion prevention, and treatment and remediation.
This chapter will deal with the latter topic.

10.2  Iron Nanotechnology

Increased soil pollution has prompted researchers and engineers to investigate


affordable and efficient strategies for environmental remediation. Contaminants are
mostly found mixed in the air, water, and soil. In particular, soil is an open, multi-
component, biogeochemical system containing solids, liquids, and gases [1]. This
means that the pollutants in the soil may be present in one or more of these phases
and a suitable technology, able to monitor, detect, and, if possible, remove the con-
taminants from all these phases is needed. In this context, nanotechnology offers a
wide range of capabilities to improve the quality of existing environment [2].
Metallic iron technology is a well-known remediation technology for organic/inor-
ganic polluted sites, able to degrade halogenated hydrocarbons [3, 4], pesticides
[5–7], and other recalcitrant organic pollutants [8–11]. The use of nanosized iron
particles allowed to reach higher pollutant removal efficiency, faster kinetics and
better mobility of the suspensions that can be directly injected in the contaminated
zone or mixed with the polluted soil in situ [12–15]. The use of iron nanoparticles
in environmental treatment field has become of great interest at the end of 1990,
when the researchers started to modify the nanoparticle’s characteristics in order to
reduce the well-known downsides, i.e., oxide shell formation and particle’s aggre-
gation [16–19]. Iron nanoparticles are characterized by a core-shell structure, with
the core made of metallic iron and the shell constituted of oxide and hydroxydes of
Fe(II) and Fe(III) [20]. nZVI may act as electron donors for the chemical reduction
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 249

of various organic and inorganic species, such as halogenated hydrocarbons,


hexavalent chromium, nitrates, chlorates; due to its high reactivity in aqueous solu-
tion, it can also be used to develop advanced oxidation processes, such as heteroge-
neous Fenton-like processes through the addition of hydrogen peroxide as main
oxidant [21–26]. During the reactions with the target pollutant, both organic and
inorganic, the nZVI core transfers electrons to the shell, where the pollutant is situ-
ated after its sorption, and the chemical reduction can occur; another typical reac-
tion is the production of hydrogen after the nZVI core oxidation that acts as an
electron acceptor [27–29]. Another important mechanism is the generation of reac-
tive oxygen species (ROS) following the nZVI core corrosion; these species are
strong oxidants and can recombine also in hydrogen peroxide, that, once reacted
with ferrous ions present onto the nZVI shell, can develop the well-known Fenton
reaction [21, 30–33].
In this paragraph, the main synthesis procedure and characteristics of iron
nanoparticles will be presented and discussed.

10.2.1  Iron Nanoparticles Synthesis

Various approaches have been proposed for the synthesis of nano-zero-valent iron
particles (nZVI), including both physical and chemical technologies belonging to
“bottom-up” and “top-down” approaches (Fig. 10.1).
Li and co-authors developed an interesting physical method to produce nZVI
without the use of any solvents or reagents [34]. In detail, the authors produced iron
nanoparticles characterized by a size up to 50 nm using a precision ball mill, starting

Fig. 10.1  typical nZVI


synthesis approaches
250 M. Stoller et al.

from micro-iron particles. The process was characterized by various advantages,


such as the possibility to synthetize the nanomaterial without the use of toxic reduc-
ing agents, solvents, or other materials/reagents different from iron. The main
downsides of the process were:
• The necessity of micro-iron, that can be purchased or produced by other equip-
ment from millimetric iron.
• The large residence time required to obtain nano-metric size (at least 8 h versus
the 15–30 min of chemical batch synthesis and less than 1 min of continuous
synthesis by a spinning disk reactor [35]).
• The significant aggregation of the final product.
Figure 10.2 displays the results obtained by Li and co-workers.
However, the most employed approach to produce nZVI is represented by the
bottom-up pathway, consisting in the direct reduction of ferrous/ferric ions in aque-
ous solution, using various reagents, such as sodium borohydride [36], hydrazine,
or natural extracts [37]. Depending on the equipment and stabilizer used, the syn-
thesis requires to be conducted under an inert atmosphere (N2 or Ar), to prevent the
rapid oxidation of the final product. For instance, fixing the reagents (iron sulfate,
sodium borohydride, and carboxy-metil-cellulose) and the solvent (water), it has
been demonstrated that the use of a spinning disk reactor, characterized by very low
micro-mixing time and residence time of ms, allows to produce nZVI without the
inert atmosphere during the synthesis [38]. Conversely, the batch-synthesis, with
classical stirred tank reactors, requires at least 15 min of residence time and the use
of an inert atmosphere during the synthesis [38]. However, both productions need
the de-oxygenation of the solvent used.
The pioneer paper by Chuan-Bao and Zhang [3], where FeCl3 was used as the
iron precursor, proposes the occurrence of the following reaction:

Fe ( H 2 O )6 + 3BH −4 + 3H 2 O → Fe 0 ↓ +3B ( OH )3 + 10.5 H 2 ↑


3+
(10.1)

and estimates that 90% of produced nanoparticles exhibit a size smaller than 100 nm
with specific surface area of 33.5 m2/g. Indeed, zero-valent iron nanoparticles can
be synthesized also by using ferrous [39] instead of ferric salts (mainly FeSO4·7H2O),

Fig. 10.2  Initial (a) and


final nanomaterial (b)
obtained by the authors
(modified from [34])
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 251

with the advantage of consuming less sodium borohydride thus limiting both costs
and hazard of the synthesis:

Fe ( H 2 O )6 + 2 BH −4 + 3H 2 O → Fe 0 ↓ +2 B ( OH )3 + 7 H 2 ↑
2+
(10.2)

In fact, a comparison of Eqs. (10.1) and (10.2) clearly indicates that the use of
ferrous ions implies a more favorable molar ratio of zero-valent iron to NaBH4 and
a lower hydrogen gas release, since the H2 produced moles passed from 10.5 to 7 per
mole of metallic iron produced. However, in spite of the stoichiometric coefficients,
either starting from ferric or ferrous ions, excess of sodium borohydride is generally
needed to enhance the reaction rate of synthesis [40] and to obtain highly reactive
nanoparticles [41]. In particular, the stoichiometric excess of borohydride used with
Fe(III) was 7.4 while with Fe(II) was 3.6 [42], even if further studies are still
required.
Usually, the nZVI synthesis requires also the introduction of a complexing agent
(stabilizer), before or after the synthesis, to reduce the mean size of the nanoparti-
cles, to modify the viscosity or, in general, the rheology of the nZVI slurry, and to
reduce the iron oxidation due to dissolved oxygen diffusion [43–46]. In particular,
among all reported stabilizers, water-soluble polysaccharides have been proved to
be the most effective stabilizer due to their low cost and environmental compatibil-
ity. It has been addressed that this kind of stabilizer not only regulates the nucleation
and particle growth during the nanoparticle formation, but also prevents agglomera-
tion of the resultant nZVI, and thus the size and reactivity of nZVI could be effec-
tively manipulated. Cellulose is a polysaccharide consisting of a series of
hydroglucose units bonded by an oxygen linkage to form a linear molecular chain
structure, while sodium carboxymethyl cellulose (CMC) is a chemical derivative of
cellulose with carboxylate groups in addition to hydroxyl groups, which may result
in strong interact between CMC and Fe nanoparticles. Usually, the CMC is added at
a molar ratio of 0.005 mol/mol with respect to Fe(II) in the iron precursor aqueous
solution, to produce a CMC–Fe(II) complex [22], and the Fe(II) complexed ions are
reduced to Fe(0) by the specific reducing agent. Conversely, another stabilizer, such
as L-cysteine, is added at the end of the Fe(0) synthesis, in higher quantity in com-
parison with CMC (0.5 mol/mol with respect to Fe(II) or Fe(III) precursor [47]). A
representation of the two stabilized slurries is schematized in Fig. 10.3.

A B

L-Cysteine

Fe

CMC

Fig. 10.3  Schematic representation of CMC and l-cysteine stabilized nZVI


252 M. Stoller et al.

As regard the equipment employed for the synthesis, nZVI have been mostly
synthetized by classical batch stirred tank reactor, equipped with mechanical stir-
rers, mainly radial impellers such as Rushton turbines [48] or with magnetic stirrers,
which usually led to the production of smaller nanoparticles but larger aggregates.
Only recently, the authors successfully produced nanosized metallic iron particles
by means of a spinning disk reactor (SDR, see Fig. 10.4), reaching better results
with respect to those obtained by batch stirred tank reactors [38]. In detail, the
authors reported a mean size of nZVI lower than 65 nm with a very narrow particle
size distribution that resulted in unimodal rather than bi-modal as that obtained by
batch stirred tank reactor.
Although these bottom-up pathways appear simple and relatively quick, without
the need of high temperature or pressure, the release of gaseous hydrogen, with
potential explosion hazard and the NaBH4 toxicity, raise concerns about the process
safety. Moreover, the high cost of NaBH4 represents a further issue and a serious
deterrent to large-scale application. Basing on these considerations, the develop-
ment of cost effective and environmentally friendly methods represents a topic of
great interest. Various green routes of nZVI synthesis include the use as the reduc-
ing and capping agent of extracts from food scrap [49] or from plants such as coffee
[37], sorghum [50], rosemary, melaleuca, eucalyptus [51], and tea [52]. Perennial
and grass plant are of great interest, since are less water and labor demanding and
can be cultivated worldwide. A size ranging from 5 to 100 nm is often reported as
well as improved stability of the nZVI, because of the capping ability of polyphe-
nols released in the extracts. The nature and concentration of polyphenols do affect
both nanoparticles size and morphology [53].
However, the feasibility of producing nZVI from natural extracts, mainly from
agro-industrial residues or directly from plants, is a controversial topic: conflicting
results regarding safety, mechanism, and effectiveness can be found in the literature.
In fact, a recent study verified that the ROS intermediates produced by Fe(II)-
catalyzed processes in water showed significant eco-toxicological impact on differ-
ent aquatic organisms due to the onset of oxidative stress [54]. Moreover, Wang has
recently proposed an alternative reaction between polyphenols and Fe(III): in detail,
it seemed that a chelated Fe(III) product was generated instead of Fe(0) [55]. Finally,

Fig. 10.4 Schematic
representation of SDR
(modified from [38])
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 253

Table 10.1  Natural extracts Natural source nZVI diameter (nm) References
and size of the produced
Tea extracts 20–120 [56]
nanoparticles
Vine leaves, grape marc 15–45 [57]
Eucalyptus leaves 20–80 [55]
Camellia sinensis 5–15 [58]
Sorghum 40–60 [59]
Dodonaea viscosa 25–30 [60]
Terminalia chebula 70–90 [61]

lower performance of green tea deriving nZVI is observed in the treatment of Cr


(VI) contaminated soils [54].
Table 10.1 summarizes the most used natural extracts and the size of the pro-
duced nanoparticles.
As reported in Table 10.1, the different extracts led to the production of quite
different nanoparticles, characterized by a size in the range 5–120  nm, but often
characterized by lower reactivity in comparison with that reported for classical-­
synthetized nZVI.

10.2.2  I ron Nanoparticle’s Characteristics and Pollutant


Removal Mechanism

The most important characteristics of nZVI are surely the high surface-to-volume
ratio, the presence of pore and defects on the external surface, the capacity to act as
an electron donor and the to act as an heterogeneous catalyst for Fenton-like reac-
tions, as well as its particular rheology when used as slurry.
The enhanced reactivity of the metallic iron at the nanoscale and therefore its
formidable performance in adsorbing/reducing a broad range of metals and organics
can be attributed to large specific surface area and reactivity, due to the presence of
active sites on its external surface. Metallic iron nanoparticles are made of a porous
shell, made initially of iron oxides such as magnetite and FeO, and a metallic core
made of Fe(0), as reported in Fig. 10.5.
The porosity of the external shell is guaranteed by the presence of iron oxide,
which are initially present mainly as Fe(II) and mixed Fe(II)/Fe(III) species, char-
acterized by high porosity and capability to sorb metals or other ionic species,
depending on the pH of the medium, on the pH of zero charge of the surface, and on
the ionic charge of the contaminant to be sorbed [62]. It has been reported by vari-
ous researchers that the pH of zero charge of nZVI is near 8, implying that to favor
the sorption and subsequent reduction of anionic species (such as chromate, sele-
nite, or anionic organic pollutants), the medium pH should be lower than 7, whereas
to improve the sorption of cations, such as metals, the pH should be basic [63]. In
the case of hydrocarbons dehalogenation, such as the degradation of PCE
254 M. Stoller et al.

Cr(III) / Fe(III) hydroxide


Fe(0) - Core

Reduction of organics and heavy metals Fe(II) / Fe (III) oxide shell

Metal sorption Cr(VI)

Fig. 10.5  Core-shell structure of Fe(0) nanoparticles

Fig. 10.6  Schematization of shrinking core mechanism (modified from [65])

(perchloroethylene), or the reduction of chromate and nitrate species, the phenom-


enological mechanism among nZVI and contaminant is a shrinking core-like reac-
tion. According to this mechanism, the contaminant needs to diffuse initially
through the liquid film around the nZVI, subsequently sorbs onto the eternal surface
and diffuses through the porosity to reach the core of the particle where the reduc-
tion reaction may occur (Fig. 10.6); the reaction product (respectively, dehaloge-
nated hydrocarbon and iron oxide, mixed Fe–Cr hydroxide, and ammonia and iron
oxide) migrate from the core to the external shell, causing the size reduction of the
core (Fe(0) consumption) and the size increase of the shell [48, 64, 65]. The nZVI
characteristics that influence the occurrence of this mechanism are mainly its spe-
cific surface area (i.e., also its mean size), its external porosity and the pH of zero
charge. Usually, the lower the size of the particle, the higher the specific surface area
at fixed shape, but at very low mean size, i.e. <10 nm, the particles result so reactive
that immediately can be oxidized by water molecule or dissolved oxygen; further-
more, even if the specific surface area increases with size reduction, the presence of
defects and thus, active sites, may decrease simultaneously, reducing the chemical
activity enhancement. However, in the absence of stabilizing agents, the nZVI
undertake many different interactions driven by magnetic attraction and Van der
Waals bonds, resulting in the formation of aggregates and possible settling.
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 255

Nanoparticles agglomeration strongly affects soil treatment, due to decreased


mobility in porous media and therefore in difficult deliverability into soils thus pre-
venting the possibility of in situ applications. To improve nanoparticles dispersion,
electrostatic and steric stabilization have been proposed. To this aim, different
approaches may be adopted [66] by modifying viscosity and surface characteristics
or by introducing solid supports. As already reported in Sect. 2.1, the surface modi-
fications of the nZVI are generally divided into pre- and post-synthesis stabilization
depending on whether stabilizers are introduced during or after the nanoparticle’s
production. Pre-synthesis methods exhibit better performance due to superior reac-
tivity and increased nanoparticles surface area [67]. A further approach is repre-
sented by immobilizing the nZVI on a support that promotes the access into porous
media. Various materials have been tested including resin [68], hydrophilized car-
bon [69] and silica fume [70]. Promising results have been obtained by using nega-
tively charged smectite clay [71]. In fact, it acts not only as a support but also as a
layer template where, due to the presence of negative charge, sparse iron ions are
linked thus resulting, after reduction, in metallic iron at a sub-nanoscale. The immo-
bilization on support produces more transportable and stable nanoparticles but
reduces the active surface, and then the reactivity. It is therefore necessary to reach
a compromise by taking into account numerous site-specific [69].
As regard the mechanism of contaminant removal by nZVI, a deeper discussion
is needed. In the soil solution-nZVI system, iron (Fe) exists in three different oxida-
tion states (0, II, and III), and soluble species may be available in a variety of aqua-
and hydroxyl complexes (e.g. Fe(OH)nn − 2, Fe(OH)nn − 3, Fe(H2O)x2+, Fe(H2O)x3+ with
n ≤ 3, and x ≤ 6). In using metals metallic species for wastewater treatment, the
main aspect of the process is that the electrochemical reduction reactions are medi-
ated by the metal surface [19, 27, 72–75]. Therefore, the dissolved pollutants must
be adsorbed onto the nZVI surface or within conductive oxide films. Alternatively,
contaminant molecules must first move on to the iron surface. This implies that the
pollutant transformation depends on the availability/accessibility of the Fe(0) sur-
face and the electronic properties of the oxide film. However, the system consists of
various solid phases, such as FeOOH, that result non-conductive and that can inhibit
the treatment process. Furthermore, according to thermodynamic considerations,
metallic iron is not necessarily the most powerful reducing agent. Therefore, con-
taminant adsorption/co-precipitation should be considered as an independent
removal mechanism in water-iron systems. The two redox systems Fe(II)/Fe(0)
(E0 = −0.44 V) and Fe(III)/Fe(II) (E0 = 0.77 V) are considered the main ones in soil
solution-iron system.; indeed, sorbed or structural Fe(II) has been showed a more
powerful reducing power with respect to Fe(0), with a E0 = −0.65 [76, 77]. Therefore,
pollutant reduction in a soil solution-iron system will not necessarily be mediated
by electrons from the bulk metal. For instance, hexavalent chromium (Cr(VI)/
Cr(III); E0 = 1.51 V) can be reduced by Fe(0) and all forms of Fe(II) [25, 78, 79],
whereas nitrate (N(V)/N(0), E0 = 0.75 V) can only be reduced by Fe(0), structural
Fe(II) and dissolved organic Fe(II). Various mechanisms may occur for the pollutant
removal through the use of nZVI: precipitation, adsorption, co-precipitation,
Fenton-like oxidation, and reduction. The distinction between precipitation, adsorp-
tion, and co-precipitation, that are usually used to remove heavy metals from
256 M. Stoller et al.

wastewater streams in adsorbing colloid flotation techniques, could result not clear.
The former usually occur when the target species, generally a metal, formed insol-
uble hydroxides in alkaline environment. Adsorption processes can occur whenever
a solid substrate surface is present. Co-precipitation can occur only if a colloid is
pre-­generated in the presence of the heavy metal ions (or generally the species) to
be removed from solution. Metal ions are then adsorbed to colloids and entrapped
in their structure while ageing [80–85]. In a soil solution-iron system at neutral pH,
iron supersaturation in the proximity of the Fe(0) surface yields to precipitation of
sorbent amorphous and crystalline iron oxyhydroxides (Fe3O4, Fe2O3, FeOOH,
Fe(OH)2, Fe(OH)3, etc.) [27–29, 86–88]. The iron oxide precipitation is not instan-
taneous; thus, some inflowing pollutants will be adsorbed on to aged iron oxyhy-
droxides; others will be adsorbed on to nascent iron oxyhydroxides and will be
entrapped in their structure while ageing (co-precipitation). Finally, local supersatu-
ration of the inflowing soil solution can yield the pollutant precipitation (at the Fe(0)
surface, within the porous oxyhydroxide matrix, or in the free pore volume). For the
pollutant chemical reduction to occur at the Fe(0) surface, the contaminant must be
transported to the surface (migration). In an undisturbed medium, the pollutant must
transfer through the liquid film that cover the nZVI particle, and must sorb onto the
shell surface, or may diffuse through the shell (that is made of porous iron species)
to reach the nZVI core. The molecular diffusion of the species (across the oxide
shell) depends on the established concentration gradients and, on the other hand, the
rate of the electrochemical processes (corrosion) depends on the species concentra-
tions at the Fe(0) surface. Therefore, there is a two-way coupling between the elec-
trochemical processes at the Fe(0) surface and processes in the adjacent solution
layer (i.e., diffusion in the oxide shell). Considering a typical continuous treatment,
the polluted soil solution moves with respect to Fe(0) surface during, for instance,
the soil-nZVI mixing process; thus, convective transport cannot be neglected.
However, for near-solid surfaces, in the oxide shell, time-averaged convection is
parallel to the surface and does not contribute to the transport of species to and from
the surface. Therefore, in the region near to the surface, no turbulence can survive
and molecular diffusion (and electro-migration) represents the predominant trans-
port mechanism (see Fig. 10.7).
The conductivity of the shell, which depends on the single chemical species pres-
ent, affects the electromigration of the species that, in turn, depends on the cross-­
sectional area available for conduction (porosity), the conductivity of the pore water,
and the complexity of the pore space (i.e., tortuosity or pore size distribution).
Basing on the scheme reported in Fig. 10.7, the transport of species in the bulk is
dominated by turbulent mixing while in the oxide shell closer to the surface is con-
trolled by molecular diffusion. In both flow fields, provided that the species are
electrically charged, electromigration plays a role, but it can be emphasized that the
role of electromigration is more significant in the oxide shell. Generally, if electro-
chemically reaction characteristic time is larger than that of molecular diffusion,
pollutant concentration variation at the Fe(0) surface will be small. If, on the con-
trary, the diffusion is slower than electrochemical reactions, a remarkable pollutant
concentration gradients can establish among the two interfacies. However,
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 257

Fig. 10.7  Double-layer flow field of pollutant transport from the bulk solution to the surface of
Fe(0) (d1 is the porous film; d2 is the diffusion sublayer; and d3 is the turbulent sublayer; it is
assumed that the turbulent field ends in the middle of d2)

considering that all solid corrosion products are of larger surface area than the Fe(0)
surface and compete with the Fe(0) surface for pollutant removal and that the Fe0
surface is (at least partly) shielded by these corrosion products, it becomes clear that
appreciable pollutant reduction at the Fe(0) surface does not occur.

10.3  Iron Nanoparticles Application for the Organic-­Polluted


Soil Remediation

The use of zero-valent iron nanoparticles for soil remediation is increasing in the
last years as a choice for treatment of hazardous and toxic substances in the soil
contaminated sites. Activities in adopting this technique are reported from USA,
Europe, and Asia. The nanoscale of these particles to boost effective subsurface
dispersion in the soil, and their large specific surface area enhances the reaction
processes to reach rapid contaminant transformation and elimination. The ongoing
innovations introduced in the field permits to achieve the nanoparticle synthesis and
production at reduced and economic feasible costs, and, as a consequence, the mar-
ket has more availability of nZVI for large scale applications. In situ experience is
collected in these and the following years, permitting the technology to become
more established. On the other hand, challenges still hold on.

10.3.1  Applications

The application of nanomaterials for remediation purposes generally involves


reduction or oxidation, in some cases assisted by embedded catalyst and biological
assistance (e.g., for the dehalorespiration where contaminants such as solvents are
biologically dechlorinated). The degradation processes lead to contaminant
258 M. Stoller et al.

degradation to simple compounds (e.g., in the treatment of chlorinated solvents) or


contaminant immobilization in insoluble forms with limited environmental acces-
sibility (e.g., in the treatment of mobile arsenic) and/or reduced toxicity (e.g., in the
conversion of chromium (VI) to chromium (III).
The application of nanoremediation requires certain operational steps and
requirements.
In the first step, the nanomaterial must be produced. In many cases, the produc-
tion of the nZVI particles is not performed on the site, but supplied. Storage of the
nZVI is a vital issue, since the product is very active and may deactivate quickly. For
this reason, sealing and capping of the product is required. The nZVI shelf-life is
not high. Once produced, the nanomaterial required to be employed as quick as pos-
sible. The supply of nZVI is mostly given as a suspension in some capping agent
and/or as a coated powder that requires activation.
The nanomaterial is then primarily deployed in situ to the surface or the subsur-
face. The reason is to activate the nanomaterial in such conditions to retrain the
active particles from direct contact to ambient air and as near as possible to the pol-
lutants, which in time settle down in the soil and migrate to the groundwater layer.
In particular, the target of the remediation may be the soil (soil decontamination),
the groundwater, or both. Depending on the final destination of the treatment, the
employment of the nZVI is different.
In case of soil remediation, the nZVI are employed by sparging and seeping. It
might be possible that the nZVI, still active, reaches the groundwater layer and
spread out by following the stream to nearby soil extending their action range. On
the contrary, for targeted groundwater remediation purposes, nanoparticles are
introduced via injection wells into the aquifer, typically requiring some form of
active pumping.
In this latter case, there are two modes of deployment by adopting in situ reme-
diation strategies: control of the source term and control of the pathway. Depending
on the site characteristics and remediation goals, these maybe applied individually
or together.
Source reduction describes the use of remediation by nZVI to directly treat a
source term, for example, a NAPL source within the saturated zone. The target of
the operation is usually to reduce the mass of the source term in order to limit the
duration of groundwater contamination. The risk of all in situ remediation tech-
niques targeting in situ source term treatments is to lead to an increased residual
sources mobility, and as a consequence, a risk of higher contamination fluxes to
groundwater. In this case the success depends on the accessibility to the source term
by the deployed treatment method, and this may limit applicability in case of adopt-
ing nZVI.
Pathway management describes the use of remediation techniques that intercept
the contaminants from the source along pathways, typically carried by groundwater
flow. In this case, since mobility of the nZVI cannot be constrained, typically the
area of treatment is confined, for example, by the use of digged permeable mem-
branes. The method can also be used to reach a source from above, that is, difficult
to reach otherwise. A main drawback was observed on the interaction between nZVI
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 259

and the confining membrane that may lead to the rupture of the latter one due to
biological based mechanisms (enzyme formation).
Another possible application of nZVI is to treat marine sediment that accumulate
inorganic contaminants and heavy metals. Remediation of contaminated sediments
has raised great interest worldwide, representing a challenge from both technical
and technological point of view. Sediment treatment strategies may involve in situ
options such as monitored natural recovery, in situ capping, and ex situ options,
such as dredging with containment and dredging with sediment treatment. A recent
study was aimed to evaluate the performances of nZVI treatment for the decontami-
nation of marine sediments polluted by heavy metals. Experimental data showed
that on marine sediment with a size less than 5 mm, which are the hardest to treat,
selective remediation resulted to be possible. Research on application of nZVI for
marine sediment decontamination is still holding on. Moreover, the challenge to
confine and follow the fate of employed nZVI in situ to marine sites is high.
A solution may be found in developing magnetic core nZVI nanoparticles that
may be recovered back after use by magnetic filters. The synthesis and production
methods of this kind of material are still under investigation. Another advantage of
the use of magnetic particles is the possibility to address them by magnets to the
point of interest for punctual treatment of sources.

10.3.2  Challenges Ahead

In year 2001 the nZVI technology was first employed in the framework of an initial
field demonstration with success. From this point on, significant progresses have
been made in the field of research, development, production, and employee of iron
nanoparticles for soil and groundwater remediation purposes. New research and
development efforts are mostly devoted to real-world applications minimizing costs
and environmental risks. Although the technique is confirmed to be feasible and
reliable, specific challenges lies ahead and are here listed:
Synthesis of nZVI: The synthesis of iron nanoparticles is the first necessary and
unavoidable step to allow this technology. Many and different nZVI synthetic routes
have been found, but overall the economics of each are still not satisfying. The chal-
lenge is to develop industrial production processes capable to produce the nanoma-
terial with high capacities, such to lead the production costs in the range of
€35–50 per kg. The actual market price is 10 times higher than the suggested value.
Moreover, the technology does not suit well to the green chemistry and circular
economy principles. Although green chemistry routes were explored in the recent
years, the relevant production costs appear to be not suitable to satisfy economic
feasibility.
Standardization of quality control and assurance: The particle size distribution,
connected to the reactivity, surface charge, and mobility of the material should be
established. The optimization of this parameter is nowadays under discussion, and
research has found different results concerning particle size. Moreover, as a
260 M. Stoller et al.

function of the soil characteristics, the value of the optimized nZVI may change,
and no reliable correlation is actually known. Other issues that will require more
detailed knowledge and standardization include optimization or customization of
the ZVI surface properties such as hydrophobicity, charge, and functional group.
These characteristics are able to allow efficient subsurface transport under site-spe-
cific conditions and for the degradation of the targeted contaminants.
Environmental chemistry and geochemistry: The deep understanding of the reac-
tion mechanisms in real-world applications is nowadays not fully developed. In
particular, in soil many interfering substances to the degradation process may be
found and distributed in the soil not homogenously. Large-scale applications may
suffer from these if the soil was not properly characterized at a punctual level. It
should be noticed that a widespread thus refined soil analysis has increased and
substantial costs. Moreover, changes in the soil of the pH, ionic strength, and com-
peting contaminants will affect the outcome of the treatment. The research in the
field of halogenated aromatic compounds, PCBs, dioxins, pesticides, and other
organic complex pollutants should be therefore increased. Remediation of metal
contaminated sites also presents interesting opportunities and challenges. Finally,
nZVI reactions with both surface and ground water, sorption, and desorption with
and through the soil and sediment, settling mechanisms, aggregation tendencies,
and the transport phenomena in porous media, all affect the treatment
performances.
Environmental impact and safety: One main drawback in the use of nanotech-
nologies is the evaluation of short- and long-term impact of nanoparticles in the
environment and their fate. Only a little amount of studies and some preliminary
tries of regulamentation were attempted in recent years. Thus far, no reports on the
ecotoxicity of nZVI in soil and water have been reported in literature. Research on
the transport of nZVI in the environment outside the application site, fate, and eco-
toxicity hazard is needed to overcome the concerns and fear the use of nanomateri-
als and to minimize the expected impacts. Moreover, iron has been found to be
included in many biological molecules and likely plays an important role in the
chemistry of living organisms. It is well known that iron is an essential element of
the blood and tissues. Iron in the body is mostly present as iron porphyrin or heme
proteins, which include hemoglobin in the blood, myoglobin, and the heme enzymes.
The challenge is to determine parallel to the ecotoxicity of the impact to human
health and life, both during application, in the short, and in the long term, when the
employed nanoparticles are out of sight and control of the application site and area.
Safety concerns are existing even at application stage, and more concern should be
given to how to handle correctly the nanomaterials. The problem is not only storage
but also to protect the health of the operator and the flora and fauna nearby.
Production and storage: once produced, nZVI are easily oxidized by the ambient.
In order to avoid oxidation and thus deactivation of the produced nZVI before their
application, proper capping should be guaranteed. Often, nZVI in capping solutions
are prepared and shipped. There is a problem of the durability, degradability, and
ecotoxicity of the capping solution. Parallel to this, polymeric coating was sug-
gested, but then some problem arise to activate the nanomaterial before use. Finally,
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 261

a recently patented application suggests to embed the nZVI particles in saline matri-
ces [89]. The adopted solution permits to have a long-term storage of the nZVI and
to handle the material easily and without hazard and to activate once sparged the
nanomaterial by dissolution of the protective matrix with water rinsing on soil,
assisting the sorption.

10.4  Conclusions

The use of nZVI in the field of soil remediation is increasing and will in the next
future. The number of contaminated sites worldwide is high and growing. The avail-
ability of a technique that reaches high efficiencies in remediation combined to an
ease of use is desired, and nZVI have all the required properties to breakthrough:
high selectivities, exhibits high performances and yield, capable to treat a wide
range of different contaminants in parallel, easy to employ (in-situ). In fact, the
technique may qualify as BAT in the next years, if all the challenges reported within
this Chapter will be faced and solved.

References

1. Sposito, G. (2008). The chemistry of soils. Oxford University Press.


2. Theron, J., Walker, J.  A., & Cloete, T.  E. (2008). Nanotechnology and water treatment:
Applications and emerging opportunities. Critical Reviews in Microbiology, 34, 43–69. https://
doi.org/10.1080/10408410701710442.
3. Chuan-Bao, W., & Zhang, W. (1997). Synthesizing nanoscale iron particles for rapid and com-
plete dechlorination of TCE and PCBs. Environmental Science & Technology, 31, 2154–2156.
https://doi.org/10.1021/ES970039C.
4. Tang, S., Wang, X., Liu, S., Yang, H., Xie, Y. F., & Yang, X. (2017). Mechanism and kinet-
ics of halogenated compound removal by metallic iron: Transport in solution, diffusion and
reduction within corrosion films. Chemosphere, 178, 119–128. https://doi.org/10.1016/J.
CHEMOSPHERE.2017.03.006.
5. O’Carroll, D., Sleep, B., Krol, M., Boparai, H., & Kocur, C. (2013). Nanoscale zero valent iron
and bimetallic particles for contaminated site remediation. Advances in Water Resources, 51,
104–122. https://doi.org/10.1016/J.ADVWATRES.2012.02.005.
6. Lu, H., Wang, J., Stoller, M., Wang, T., Bao, Y., & Hao, H. (2016). An overview of nanomateri-
als for water and wastewater treatment. Advances in Materials Science and Engineering, 2016,
1–10. https://doi.org/10.1155/2016/4964828.
7. Tratnyek, P. G., & Johnson, R. L. (2006). Nanotechnologies for environmental cleanup. Nano
Today, 1, 44–48. https://doi.org/10.1016/S1748-­0132(06)70048-­2.
8. Takayanagi, A., Kobayashi, M., & Kawase, Y. (2017). Removal of anionic surfactant sodium
dodecyl benzene sulfonate (SDBS) from wastewaters by zero-valent iron (ZVI): Predominant
removal mechanism for effective SDBS removal. Environmental Science and Pollution
Research, 24, 8087–8097. https://doi.org/10.1007/s11356-­017-­8493-­8.
9. Li, B., & Zhu, J. (2014). Removal of p-chloronitrobenzene from groundwater: Effectiveness
and degradation mechanism of a heterogeneous nanoparticulate zero-valent iron (NZVI)-
induced Fenton process. Chemical Engineering Journal, 255, 225–232.
262 M. Stoller et al.

10. Epolito, W. J., Yang, H., Bottomley, L. A., & Pavlostathis, S. G. (2008). Kinetics of zero-valent
iron reductive transformation of the anthraquinone dye Reactive Blue 4. Journal of Hazardous
Materials, 160, 594–600. https://doi.org/10.1016/J.JHAZMAT.2008.03.033.
11. Taha, M. R., & Ibrahim, A. H. (2014). Characterization of nano zero-valent iron (nZVI) and
its application in sono-Fenton process to remove COD in palm oil mill effluent. Journal of
Environmental Chemical Engineering, 2, 1–8. https://doi.org/10.1016/J.JECE.2013.11.021.
12. Zha, S., Cheng, Y., Gao, Y., Chen, Z., Megharaj, M., & Naidu, R. (2014). Nanoscale zero-valent
iron as a catalyst for heterogeneous Fenton oxidation of amoxicillin. Chemical Engineering
Journal, 255, 141–148. https://doi.org/10.1016/J.CEJ.2014.06.057.
13. Bae, Y., & Park, J.-W. (2009). TCE reduction modeling in soil column: Effect of zero-valent
iron, ferrous iron, and iron-reducing bacteria. Desalination and Water Treatment, 4, 229–232.
https://doi.org/10.5004/dwt.2009.487.
14. Yan, W., Herzing, A. A., Kiely, C. J., & Zhang, W. (2010). Nanoscale zero-valent iron (nZVI):
Aspects of the core-shell structure and reactions with inorganic species in water. Journal of
Contaminant Hydrology, 118, 96–104. https://doi.org/10.1016/J.JCONHYD.2010.09.003.
15. Mpouras, T., Panagiotakis, I., Dermatas, D., & Chrysochoou, M. (2014). Nano-zero-valent
iron: An emerging technology for contaminated site remediation. In Geo-Congress 2014
Tech. Pap (pp.  2206–2215). Reston, VA: American Society of Civil Engineers. https://doi.
org/10.1061/9780784413272.215.
16. Fu, F., Dionysiou, D. D., & Liu, H. (2014). The use of zero-valent iron for groundwater reme-
diation and wastewater treatment: A review. Journal of Hazardous Materials, 267, 194–205.
https://doi.org/10.1016/j.jhazmat.2013.12.062.
17. Li, L., Fan, M., Brown, R.  C., Van Leeuwen, J., Wang, J., Wang, W., Song, Y., & Zhang,
P. (2006). Synthesis, properties, and environmental applications of nanoscale iron-based
materials: A review. Critical Reviews in Environmental Science and Technology, 36, 405–431.
https://doi.org/10.1080/10643380600620387.
18. Nurmi, J.  T., Tratnyek, P.  G., Sarathy, V., Baer, D.  R., Amonette, J.  E., Pecher, K., Wang,
C., Linehan, J. C., Matson, D. W., Penn, R. L., & Driessen, M. D. (2005). Characterization
and properties of metallic iron nanoparticles: Spectroscopy, electrochemistry, and kinetics.
Environmental Science & Technology, 39, 1221–1230. https://doi.org/10.1021/es049190u.
19. Li, X., Elliott, D. W., & Zhang, W. (2006). Zero-valent iron nanoparticles for abatement of
environmental pollutants: Materials and engineering aspects. Critical Reviews in Solid State
and Materials Sciences, 31, 111–122. https://doi.org/10.1080/10408430601057611.
20. Cirtiu, C. M., Raychoudhury, T., Ghoshal, S., & Moores, A. (2011). Systematic comparison
of the size, surface characteristics and colloidal stability of zero valent iron nanoparticles
pre- and post-grafted with common polymers. Colloids and Surfaces A: Physicochemical and
Engineering Aspects, 390, 95–104. https://doi.org/10.1016/J.COLSURFA.2011.09.011.
21. Vilardi, G., Sebastiani, D., Miliziano, S., Verdone, N., & Di Palma, L. (2018). Heterogeneous
nZVI-induced Fenton oxidation process to enhance biodegradability of excavation by-­products.
Chemical Engineering Journal, 335, 309–320. https://doi.org/10.1016/j.cej.2017.10.152.
22. Vilardi, G., Di Palma, L., & Verdone, N. (2018). On the critical use of zero valent iron nanopar-
ticles and Fenton processes for the treatment of tannery wastewater. Journal of Water Process
Engineering, 22C, 109–122.
23. Muradova, G. G., Gadjieva, S. R., Di Palma, L., & Vilardi, G. (2016). Nitrates removal by
bimetallic nanoparticles in water. Chemical Engineering Transactions, 47, 205–210. https://
doi.org/10.3303/CET1647035.
24. Vilardi, G. (2018). Bimetallic nZVI-induced chemical denitrification modelling using the
shrinking core model. Chemical Engineering Transactions, 70, 235–241.
25. Vilardi, G., Mpouras, T., Dermatas, D., Verdone, N., Polydera, A., & Di Palma, L. (2018).
Nanomaterials application for heavy metals recovery from polluted water: The combination
of nano zero-valent iron and carbon nanotubes. Competitive adsorption non-linear modeling.
Chemosphere, 201, 716–729. https://doi.org/10.1016/j.chemosphere.2018.03.032.
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 263

26. Vilardi, G., Verdone, N., & Di Palma, L. (2017). The influence of nitrate on the reduction of
hexavalent chromium by zero-valent iron nanoparticles in polluted wastewater. Desalination
and Water Treatment, 86, 252–258. https://doi.org/10.5004/dwt.2017.20710.
27. Keenan, C. R., & Sedlak, D. L. (2008). Factors affecting the yield of oxidants from the reac-
tion of nanoparticulate zero-valent iron and oxygen. Environmental Science & Technology, 42,
1262–1267. https://doi.org/10.1021/es7025664.
28. He, D., Ma, J., Collins, R. N., & Waite, T. D. (2016). Effect of structural transformation of
nanoparticulate zero-valent iron on generation of reactive oxygen species. Environmental
Science & Technology, 50, 3820–3828. https://doi.org/10.1021/acs.est.5b04988.
29. Liu, A., Liu, J., & Zhang, W. X. (2015). Transformation and composition evolution of nanoscale
zero valent iron (nZVI) synthesized by borohydride reduction in static water. Chemosphere,
119, 1068–1074. https://doi.org/10.1016/j.chemosphere.2014.09.026.
30. Kuang, Y., Wang, Q., Chen, Z., Megharaj, M., & Naidu, R. (2013). Heterogeneous Fenton-­
like oxidation of monochlorobenzene using green synthesis of iron nanoparticles. Journal of
Colloid and Interface Science, 410, 67–73. https://doi.org/10.1016/j.jcis.2013.08.020.
31. Xu, L., & Wang, J. (2011). A heterogeneous Fenton-like system with nanoparticulate zero-­
valent iron for removal of 4-chloro-3-methyl phenol. Journal of Hazardous Materials, 186,
256–264. https://doi.org/10.1016/j.jhazmat.2010.10.116.
32. Vilardi, G., Ochando-Pulido, J. M., Stoller, M., Verdone, N., & Di Palma, L. (2018). Fenton
oxidation and chromium recovery from tannery wastewater by means of iron-based coated
biomass as heterogeneous catalyst in fixed-bed columns. Chemical Engineering Journal,
351, 1–11.
33. Vilardi, G., Rodriguez-Rodriguez, J., Ochando Pulido, J.  M., Verdone, N., Martinez-Ferez,
A., & Di Palma, L. (2018). Large laboratory-plant application for the treatment of a Tannery
wastewater by Fenton oxidation: Fe(II) and nZVI catalysts comparison and kinetic modelling.
Process Safety and Environment Protection, 117, 629–638.
34. Li, S., Yan, W., Zhang, W., Nielsen, M. M., Morup, S., Pecher, K., Wang, C. M., Linehan, J. C.,
Matson, D. W., Penn, R. L., & Driessen, M. D. (2009). Solvent-free production of nanoscale
zero-valent iron (nZVI) with precision milling. Green Chemistry, 11, 1618–1626. https://doi.
org/10.1039/b913056j.
35. Stoller, M., Vilardi, G., Di Palma, L., Chianese, A., & Morganti, P. (2017). Process intensi-
fication techniques for the production of nanoparticles for the cosmetic and pharmaceutical
industry. Journal of Applied Cosmetology, 35, 53–59.
36. Di Palma, L., Verdone, N., & Vilardi, G. (2018). Kinetic modeling of Cr(VI) reduction by
nZVI in soil: The influence of organic matter and manganese oxide. Bulletin of Environmental
Contamination and Toxicology, 101, 692–697.
37. Kozma, G., Rónavári, A., Kónya, Z., & Kukovecz, Á. (2016). Environmentally benign synthe-
sis methods of zero-valent iron nanoparticles. ACS Sustainable Chemistry & Engineering, 4,
291–297. https://doi.org/10.1021/acssuschemeng.5b01185.
38. Vilardi, G., Stoller, M., Verdone, N., & Di Palma, L. (2017). Production of nano zero valent
iron particles by means of a spinning disk reactor. Chemical Engineering Transactions, 57,
751–756. https://doi.org/10.1007/s00128-­016-­1865-­9.
39. Glavee, G. N., Klabunde, J. K., Sorensen, J. M. C., & Hadjipanayisld, G. C. (1995). Chemistry
of borohydride reduction of Iron(II) and Iron(II) ions in aqueous and nonaqueous media.
Formation of Nanoscale Fe, FeB, and Fe2B powders. Inorganic Chemistry, 34, 28–35. http://
pubs.acs.org.ezproxy.uniroma1.it/doi/pdf/10.1021/ic00105a009 (accessed August 22, 2017).
40. Sun, Y.-P., Li, X., Cao, J., Zhang, W., & Wang, H. P. (2006). Characterization of zero-valent
iron nanoparticles. Advances in Colloid and Interface Science, 120, 47–56. https://doi.
org/10.1016/j.cis.2006.03.001.
41. Hwang, Y.-H., Kim, D.-G., & Shin, H.-S. (2011). Effects of synthesis conditions on the char-
acteristics and reactivity of nano scale zero valent iron. Applied Catalysis B: Environmental,
105, 144–150. https://doi.org/10.1016/J.APCATB.2011.04.005.
264 M. Stoller et al.

42. Zhang, W., & Elliott, D. W. (2006). Applications of iron nanoparticles for groundwater reme-
diation. Remediation Journal, 16, 7–21. https://doi.org/10.1002/rem.20078.
43. Jia, Z., Shu, Y., Huang, R., Liu, J., & Liu, L. (2018). Enhanced reactivity of nZVI embedded into
supermacroporous cryogels for highly efficient Cr(VI) and total Cr removal from aqueous solu-
tion. Chemosphere, 199, 232–242. https://doi.org/10.1016/J.CHEMOSPHERE.2018.02.021.
44. Taha, M. R., & Ibrahim, A. H. (2014). Characterization of nano zero-valent iron (nZVI) and
its application in sono-Fenton process to remove COD in palm oil mill effluent. Journal of
Environmental Chemical Engineering, 2, 1–8. https://doi.org/10.1016/J.JECE.2013.11.021.
45. Sun, Y.-P., Li, X.-Q., Zhang, W.-X., & Wang, H. P. (2007). A method for the preparation of
stable dispersion of zero-valent iron nanoparticles. Colloids and Surfaces A: Physicochemical
and Engineering Aspects, 308, 60–66. https://doi.org/10.1016/J.COLSURFA.2007.05.029.
46. Wang, X., Le, L., Wang, A., Liu, H., Ma, J., & Li, M. (2016). Comparative study on proper-
ties, mechanisms of anionic dispersant modified nano zero-valent iron for removal of Cr(VI).
Journal of the Taiwan Institute of Chemical Engineers, 66, 115–125. https://doi.org/10.1016/J.
JTICE.2016.05.049.
47. Bagbi, Y., Sarswat, A., Tiwari, S., Mohan, D., Pandey, A., & Solanki, P. R. (2017). Synthesis
of l-cysteine stabilized zero-valent iron (nZVI) nanoparticles for lead remediation from
water. Environmental Nanotechnology, Monitoring and Management, 7, 34–45. https://doi.
org/10.1016/J.ENMM.2016.11.008.
48. Vilardi, G., Di Palma, L., & Verdone, N. (2019). A physical-based interpretation of mecha-
nism and kinetics of Cr(VI) reduction in aqueous solution by zero-valent iron nanoparticles.
Chemosphere, 220, 590–599.
49. Wei, Y., Fang, Z., Zheng, L., Tan, L., & Tsang, E. P. (2016). Green synthesis of Fe nanopar-
ticles using Citrus maxima peels aqueous extracts. Materials Letters, 185, 384–386. https://
doi.org/10.1016/j.matlet.2016.09.029.
50. Njagi, E. C., Huang, H., Stafford, L., Genuino, H., Galindo, H. M., Collins, J. B., Hoag, G. E.,
& Suib, S. L. (2011). Biosynthesis of iron and silver nanoparticles at room temperature using
aqueous sorghum bran extracts. Langmuir, 27, 264–271. https://doi.org/10.1021/la103190n.
51. Wang, Z., Fang, C., & Megharaj, M. (2014). Characterization of iron–polyphenol nanoparti-
cles synthesized by three plant extracts and their fenton oxidation of azo dye. ACS Sustainable
Chemistry & Engineering, 2, 1022–1025. https://doi.org/10.1021/sc500021n.
52. Hoag, G.  E., Collins, J.  B., Holcomb, J.  L., Hoag, J.  R., Nadagouda, M.  N., & Varma,
R.  S. (2009). Degradation of bromothymol blue by ‘greener’ nano-scale zero-valent iron
synthesized using tea polyphenols. Journal of Materials Chemistry, 19, 8671. https://doi.
org/10.1039/b909148c.
53. Markova, Z., Novak, P., Kaslik, J., Plachtova, P., Brazdova, M., Jancula, D., Siskova, K. M.,
Machala, L., Marsalek, B., Zboril, R., & Varma, R. (2014). Iron(II,III)–Polyphenol com-
plex nanoparticles derived from Green Tea with remarkable ecotoxicological impact. ACS
Sustainable Chemistry & Engineering, 2, 1674–1680. https://doi.org/10.1021/sc5001435.
54. Wang, Y., Fang, Z., Liang, B., & Tsang, E.  P. (2014). Remediation of hexavalent chro-
mium contaminated soil by stabilized nanoscale zero-valent iron prepared from steel pick-
ling waste liquor. Chemical Engineering Journal, 247, 283–290. https://doi.org/10.1016/j.
cej.2014.03.011.
55. Wang, T., Jin, X., Chen, Z., Megharaj, M., & Naidu, R. (2014). Green synthesis of Fe nanopar-
ticles using eucalyptus leaf extracts for treatment of eutrophic wastewater. Science of the Total
Environment, 466–467, 210–213. https://doi.org/10.1016/J.SCITOTENV.2013.07.022.
56. Huang, L., Weng, X., Chen, Z., Megharaj, M., & Naidu, R. (2014). Green synthesis of iron
nanoparticles by various tea extracts: Comparative study of the reactivity. Spectrochimica Acta
Part A: Molecular and Biomolecular Spectroscopy, 130, 295–301. https://doi.org/10.1016/J.
SAA.2014.04.037.
57. Machado, S., Stawiński, W., Slonina, P., Pinto, A. R., Grosso, J. P., Nouws, H. P. A., Albergaria,
J. T., & Delerue-Matos, C. (2013). Application of green zero-valent iron nanoparticles to the
10  Soil Remediation by Zero-Valent Iron Nanoparticles for Organic Pollutant… 265

remediation of soils contaminated with ibuprofen. Science of the Total Environment, 461–462,
323–329. https://doi.org/10.1016/J.SCITOTENV.2013.05.016.
58. Hoag, G.  E., Collins, J.  B., Holcomb, J.  L., Hoag, J.  R., Nadagouda, M.  N., & Varma,
R. S. (2009). Degradation of bromothymol blue by ‘greener’ nano-scale zero-valent iron syn-
thesized using tea polyphenols. Journal of Materials Chemistry, 19, 8671–8677. https://doi.
org/10.1039/b909148c.
59. Njagi, E. C., Huang, H., Stafford, L., Genuino, H., Galindo, H. M., Collins, J. B., Hoag, G. E.,
& Suib, S. L. (2011). Biosynthesis of iron and silver nanoparticles at room temperature using
aqueous sorghum bran extracts. Langmuir, 27, 264–271. https://doi.org/10.1021/la103190n.
60. Kiruba Daniel, S. C. G., Vinothini, G., Subramanian, N., Nehru, K., & Sivakumar, M. (2013).
Biosynthesis of Cu, ZVI, and Ag nanoparticles using Dodonaea viscosa extract for antibac-
terial activity against human pathogens. Journal of Nanoparticle Research, 15, 1319–1328.
https://doi.org/10.1007/s11051-­012-­1319-­1.
61. Mohan Kumar, K., Mandal, B. K., Siva Kumar, K., Sreedhara Reddy, P., & Sreedhar, B. (2013).
Biobased green method to synthesise palladium and iron nanoparticles using Terminalia che-
bula aqueous extract. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy,
102, 128–133. https://doi.org/10.1016/J.SAA.2012.10.015.
62. Ling, L., Huang, X., Li, M., & Zhang, W. (2017). Mapping the reactions in a single zero-­
valent iron nanoparticle. Environmental Science & Technology, 51, 14293–14300. https://doi.
org/10.1021/acs.est.7b02233.
63. Ji, Y. (2014). Ions removal by iron nanoparticles: A study on solid–water interface with zeta
potential. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 444, 1–8.
https://doi.org/10.1016/J.COLSURFA.2013.12.031.
64. Tsakiroglou, C. D., Hajdu, K., Terzi, K., Aggelopoulos, C., & Theodoropoulou, M. (2017).
A statistical shrinking core model to estimate the overall dechlorination rate of PCE by an
assemblage of zero-valent iron nanoparticles. Chemical Engineering Science, 167, 191–203.
https://doi.org/10.1016/J.CES.2017.04.007.
65. Vilardi, G. (2019). Mathematical modelling of simultaneous nitrate and dissolved oxygen
reduction by Cu-nZVI using a bi-component shrinking core model. Powder Technology, 343,
613–618. https://doi.org/10.1016/J.POWTEC.2018.11.082.
66. Saleh, N., Sirk, K., Liu, Y., Phenrat, T., Dufour, B., Matyjaszewski, K., Tilton, R. D., & Lowry,
G. V. (2007). Surface modifications enhance nanoiron transport and NAPL targeting in satu-
rated porous media. Environmental Engineering Science, 24, 45–57. https://doi.org/10.1089/
ees.2007.24.45.
67. Franco, D. V., Da Silva, L. M., & Jardim, W. F. (2009). Reduction of hexavalent chromium in
soil and ground water using zero-valent iron under batch and semi-batch conditions. Water, Air,
and Soil Pollution, 197, 49–60. https://doi.org/10.1007/s11270-­008-­9790-­0.
68. Ponder, S. M., Darab, J. G., & Mallouk, T. E. (2000). Remediation of Cr(VI) and Pb(II) aqueous
solutions using supported, nanoscale zero-valent iron. Environmental Science & Technology,
34, 2564–2569. https://doi.org/10.1021/ES9911420.
69. Bettina, S., Hydutsky, B.  W., Blough, J.  L., & Mallouk, T.  E. (2004). Delivery vehicles
for zerovalent metal nanoparticles in soil and groundwater. Chemistry of Materials, 16,
2187–2193. https://doi.org/10.1021/CM0218108.
70. Li, Y., Jin, Z., Li, T., & Li, S. (2011). Removal of hexavalent chromium in soil and ground-
water by supported nano zero-valent iron on silica fume. Water Science and Technology, 63.
http://wst.iwaponline.com/content/63/12/2781 (accessed August 22, 2017).
71. Gu, C., Li, H., Teppen, B. J., & Boyd, S. A. (2013). Highly reactive subnano-sized zero-valent
iron synthesized on smectite clay templates. Functions of Natural Organic Matter in Changing
Environment, 789–792. https://doi.org/10.1007/978-­94-­007-­5634-­2_143.
72. Di Palma, L., Gueye, M. T., & Petrucci, E. (2015). Hexavalent chromium reduction in contam-
inated soil: A comparison between ferrous sulphate and nanoscale zero-valent iron. Journal of
Hazardous Materials, 281, 70–76. https://doi.org/10.1016/j.jhazmat.2014.07.058.
266 M. Stoller et al.

73. Zhao, X., Liu, W., Cai, Z., Han, B., Qian, T., & Zhao, D. (2016). An overview of preparation
and applications of stabilized zero-valent iron nanoparticles for soil and groundwater remedia-
tion. Water Research, 100, 245–266. https://doi.org/10.1016/j.watres.2016.05.019.
74. Xue, X., Lv, X., Jiang, G., Baig, S. A., & Xu, X. (2016). Influence of environmental factors on
hexavalent chromium removal from aqueous solutions by nano-adsorbent composites. Clean:
Soil, Air Water, 218, 55–64. https://doi.org/10.1002/clen.201400448.
75. Watts, M. P., Coker, V. S., Parry, S. A., Pattrick, R. A. D., Thomas, R. A. P., Kalin, R., & Lloyd,
J. R. (2015). Biogenic nano-magnetite and nano-zero valent iron treatment of alkaline Cr(VI)
leachate and chromite ore processing residue. Applied Geochemistry, 54, 27–42. https://doi.
org/10.1016/j.apgeochem.2014.12.001.
76. Noubactep, C. (2015). Metallic iron for environmental remediation: A review of reviews.
Water Research, 85, 114–123. https://doi.org/10.1016/j.watres.2015.08.023.
77. Mwakabona, H. T., Ndé-Tchoupé, A. I., Njau, K. N., Noubactep, C., & Wydra, K. D. (2017).
Metallic iron for safe drinking water provision: Considering a lost knowledge. Water Research,
117, 127–142. https://doi.org/10.1016/J.WATRES.2017.03.001.
78. Xiao-qin, L., Jiasheng, C., & Zhang, W. (2008). Stoichiometry of Cr(VI) immobilization
using Nanoscale Zerovalent Iron (nZVI): A study with high-resolution X-Ray Photoelectron
Spectroscopy (HR-XPS). Industrial and Engineering Chemistry Research, 47, 2131–2139.
https://doi.org/10.1021/IE061655X.
79. Bartlett, R.  J., & Kimble, J.  M. (1976). Behavior of chromium in soils: II.  Hexavalent
form. Journal of Environmental Quality, 5, 383. https://doi.org/10.2134/jeq197
6.00472425000500040010x.
80. Sharma, R.  K., & Agrawal, M. (2005). Biological effects of heavy metals: An overview.
Journal of Environmental Biology, 26, 301–313. ISSN: 0254–8704.
81. Khan, A., Khan, S., Khan, M. A., Qamar, Z., & Waqas, M. (2015). The uptake and bioaccumu-
lation of heavy metals by food plants, their effects on plants nutrients, and associated health
risk: A review. Environmental Science and Pollution Research, 22, 13772–13799. https://doi.
org/10.1007/s11356-­015-­4881-­0.
82. Tchounwou, P. B., Yedjou, C. G., Patlolla, A. K., & Sutton, D. J. (2012). Heavy metals toxicity
and the environment. Molecular, Clinical and Environmental Toxicology, 133–164. https://doi.
org/10.1007/978-­3-­7643-­8340-­4_6.
83. Akpor, O.  B. (2014). Heavy metal pollutants in wastewater effluents: Sources, effects and
remediation. Advances in Bioscience and Bioengineering, 2, 37. https://doi.org/10.11648/j.
abb.20140204.11.
84. Bolan, N., Kunhikrishnan, A., Thangarajan, R., Kumpiene, J., Park, J., Makino, T., Kirkham,
M.  B., & Scheckel, K. (2014). Remediation of heavy metal(loid)s contaminated soils  – To
mobilize or to immobilize? Journal of Hazardous Materials, 266, 141–166. https://doi.
org/10.1016/j.jhazmat.2013.12.018.
85. Vilardi, G., Di Palma, L., & Verdone, N. (2018). Heavy metals adsorption by banana peels
micro-powder. Equilibrium modeling by non-linear models. Chinese Journal of Chemical
Engineer, 26, 455–464. https://doi.org/10.1016/j.cjche.2017.06.026.
86. Sheng, G., Alsaedi, A., Shammakh, W., Monaquel, S., Sheng, J., Wang, X., Li, H., & Huang,
Y. (2016). Enhanced sequestration of selenite in water by nanoscale zero valent iron immobili-
zation on carbon nanotubes by a combined batch, XPS and XAFS investigation. Carbon N. Y.,
99, 123–130. https://doi.org/10.1016/j.carbon.2015.12.013.
87. Kanel, S.  R., Manning, B., Charlet, L., & Choi, H. (2005). Removal of Arsenic(III) from
groundwater by nanoscale zero-valent iron. Environmental Science & Technology, 39,
1291–1298. https://doi.org/10.1021/ES048991U.
88. Song, Y. W., Shan, D. Y., & Han, E. H. (2008). Electrodeposition of hydroxyapatite coating
on AZ91D magnesium alloy for biomaterial application. Materials Letters, 62, 3276–3279.
https://doi.org/10.1016/j.matlet.2008.02.048.
89. Patent No. PCT/IB2019/052683.
Chapter 11
Black TiO2: An Emerging Photocatalyst
and Its Applications

P. Anil Kumar Reddy, P. Venkata Laxma Reddy,
and S. V. Prabhakar Vattikuti

11.1  Introduction

The TiO2-­based photocatalysis has a wide range of applications in the field of the
environmental research [1–3]. In the recent years, titania-­based materials are exten-
sively investigated immensely in for the treatment of the pollutants, hydrogen pro-
duction, disinfection, and so on [4–6]. Although various other semiconductors as
well possess these catalysis properties, the titania has proven to be more advanta-
geous due to their low toxicity and economic feasibility [7, 8]. In general, the basic
mechanism of TiO2 photocatalysis involves the generation of the electrons and holes
upon its surface upon the absorption of the light energy higher than its bandgap. It
is desired that these photogenerated charges must be separated and migrate to the
semiconductor surface rather than undergoing the recombination [9]. In addition,
the energy that is necessary for the excitation of the charges is also desired to be in
visible region rather than the artificial illumination source such as UV and so on.
On this note, there are various modifications that have been proposed in the
recent years to improve its absorption in visible light region as well as to decrease
the recombination [8, 10]. The primary and most widely employed modification
was the doping of the metal ions. In this case, the parent titania material was doped
with an appropriate amount of metal such as transition metal ions [3, 4, 11, 12].
Consequentially, the metal doping facilitates the formation of the electron capture
centers in the bandgap. Nonetheless, such treatment can alter the crystallinity to

P. A. K. Reddy (*)
School of Mechanical and Nuclear Engineering, Ulsan National Institute of Science and
Technology (UNIST), Ulsan, South Korea
P. V. L. Reddy
Program in Environmental Science and Engineering, University of Texas El Paso,
El Paso, TX, USA
S. V. P. Vattikuti
School of Mechanical Engineering, Yeungnam University, Gyeongsan, South Korea

© Springer Nature Switzerland AG 2021 267


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_11
268 P. A. K. Reddy et al.

minimize electron/hole recombination. The doping was also extended to the non-­­
metal such as carbon, sulfur, and so on [13, 14]. These non-­metal doping has led to
the upward movement of valence band as well as the formation of oxygen-­deficient
sites. These above aforementioned strategies are usually very successful in the
decreased recombination. On the other end, there are different kinds of the modifi-
cations such as dye sensitization that are more successful in improving the visible
light absorption of the titania materials [15, 16]. Among various such modifications
to enhance the visible light absorption, the recently discovered “black titania” has
attracted huge attention for its unique properties [17]. Considering its emerging
importance, we tried to briefly outline the basic characteristics of the black titania
and its efficiency in the various applications such as treatment of aqueous pollut-
ants, hydrogen production, and CO2 reforming.

11.2  Black TiO2 Preparation Methods

Several methods have been employed for the synthesis of black TiO2 toward various
functional applications (Fig. 11.1). Hydrogenation treatment in hydrogen-­nitrogen
or hydrogen-­argon gas mixtures, high-­or low-­pressure hydrogen treatment, electro-
chemical reduction, hydrogen plasma treatment, chemical reduction, or hydroxyl-
ation pulsed laser ablation are the primary methods developed to synthesize black
TiO2 [18–20].

Fig. 11.1  Classification of various preparation methods of black TiO2


11  Black TiO2: An Emerging Photocatalyst and Its Applications 269

11.2.1  Thermal Treatment

Black TiO2 is synthesized through a wide range of hydrogenation-­based approaches.


It is perceived that the hydrogenation of TiO2 by thermal treatment is simple; how-
ever, this reaction is not modest as we think. A general thinking is that the reduction
of TiO2 transforms the Ti4+ into Ti3+ and other reduction states followed by the
changes in the structure, physical, and chemical properties. Chemical properties of
the starting material, the reaction temperature, pressure, and concentration of hydro-
gen play an important role in the synthesis of black TiO2. Moreover, the size, shape,
surface morphology, and defects also contribute to the variation in the properties of
final product. Thus, these variables make the reaction highly complicated.
Experimentally, black TiO2 is prepared by adopting various conditions to fine tune
the chemical/physical properties and performance of the black TiO2 toward differ-
ent needs [19]. The following are some methods to produce black TiO2 by thermal
treatment.

High-Pressure Pure Hydrogen Treatment

High-­pressure hydrogen treatment is the most studied and prominent method to


prepare black TiO2. Apparently, the hydrogenation condition played a critical role in
the final color and optical properties of the black TiO2. In this method the ultrahigh
pure hydrogen is used with a gas pressure in the range of 10–70 bar. Accordingly,
the reaction temperature and reaction time are also varied in the range of 200–500 °C
and 1 h–20 days, respectively. The reaction time is irreversibly proportional to the
reaction temperature and pressure of the gas. Short reaction time is required for
high-­temperature and high-­pressure conditions and vice versa. Qiu et  al. synthe-
sized blue hydrogenated rutile TiO2 nanoparticles by heating rutile TiO2 nanoparti-
cles under 40 bar hydrogen pressure at 450 °C for 1 h [21]. Similarly, Liu et al.
observed light blue colored TiO2 under pressurized H2 at 20 bar and 500 °C reaction
temperature for 1 h [22]. However, the reaction time was prolonged to several hours
and days when the reaction temperature and pressure is low. Chen et al. described
the black TiO2 preparation at 20 bar pure hydrogen at about 200 °C for 5 days [23].
Lu et al. systematically studied the room temperature hydrogenation of white P25
TiO2 in high-­pressure hydrogen atmosphere at about 35 bar for a period of 20 days
[24]. They observed the color changes of the hydrogenated TiO2 over time. The
white color of P25 TiO2 was changed to pale yellow initially (3 days) that turned
darker with time, and a gray color was observed after 15 days. The black TiO2 was
only observed after 15–20 days reaction time (Fig. 11.2) [24]. In an experimental
study carried out by Liu et al., the black titania exhibited TiO2@TiO2−x core-­shell
structure which is attributed to hydrogenation. These black titania samples were
observed to have mid-­gap electronic states which led to the overall improvement of
visible light absorption of the sample as well as the effective separation of the
charges [25].
270 P. A. K. Reddy et al.

Fig. 11.2 Pictures
showing the effect of
hydrogenation time on the
color of hydrogenated TiO2
[24]. Reprinted with
permission from [24].
Copyright (2014) The
Royal Society of
Chemistry

Ambient or Low-Pressure Pure Hydrogen Treatment

As mentioned above, many reports claimed the black TiO2 at high hydrogen pres-
sure conditions. Contrarily, some researchers have successfully prepared black TiO2
at low and ambient hydrogen pressure. However, the low-­pressure conditions were
corroborated by the high reaction temperatures. The minimum temperature was
observed to be 500 °C to synthesize black TiO2 at low H2 pressure conditions. Liu
et al. treated anatase TiO2 nanoparticles under pure H2 flow with a general grade
from 20 to 700 °C for 2 h [26]. Yu et al. accomplished hydrogenation of anatase
TiO2 under low H2 flow of 50 sccm at atmospheric pressure in the range of
500–700 °C reaction temperature [27]. They observe the color change from white to
blue and then gray depending on the reaction temperature and time as shown in the
Fig. 11.3 [27]. Similarly, Wang et al. prepared hydrogenated rutile TiO2 nanowire
arrays using ultrahigh pure hydrogen at ambient pressure and temperatures ranging
from 200 to 550 °C for 3 h. Based on the annealing temperature, they observed the
white colored TiO2 turned into different colors, i.e., yellow green (at 350 °C) and
black (450 °C) as shown in Fig. 11.4 [28]. In another approach, Naldoni et al. con-
verted the amorphous TiO2 into black TiO2 nanoparticles by heating in hydrogen
stream at 500 °C for 1 h. When these hydrogenated samples are exposed to air or
cooled slowly, they develop gray or pale blue-­colored samples with reduced crystal-
linity and unmodified absorption. However, the samples retained their color and
crystallinity when the samples were subjected to rapid cooling in inert atmosphere
after the hydrogenation [29].
11  Black TiO2: An Emerging Photocatalyst and Its Applications 271

Fig. 11.3  Photos of


hydrogenated anatase TiO2
nanosheets synthesized in
H2 environment at
500–700 °C temperature
[27]. Reprinted with
permission from [27].
Copyright (2013)
American Chemical
Society

Fig. 11.4  Photographs of


rutile TiO2 nanowires
treated in ultrahigh pure
hydrogen at different
temperatures for 3 h [28].
Reprinted with permission
from [28]. Copyright
(2011) American Chemical
Society

Ambient Hydrogen–Argon Treatment

Alternative method to produce black TiO2 is using hydrogen–argon treatment. Apart


from pure hydrogen treatment, a mixture of hydrogen and noble gas (generally
argon) was also considered to produce black TiO2 at ambient reaction conditions. In
this method, the ratio of hydrogen to argon typically fixed at 5 v/v% however can be
increased up to 25 v/v%. The operated temperatures for hydrogenation vary between
400 and 600 °C. The advantage of this approach lies in the short reaction time which
is typically 1–5 h. Shin et al. hydrogenated the commercial TiO2 nanoparticles at
450  °C for 1–7  h under a constant flow of 5  v/v% of H2/Ar gas mixture. They
observed the color of pristine TiO2 changed from pure white to light yellowish (1 h)
followed by the dark yellow (7 h) [30]. In another work, Danon et al. performed
hydrogenation of titanite in 5 v/v% H2/Ar mixture at 350 °C for 3 h. Interestingly,
they found that the color of the final product depends on the reactor in which the
hydrogenation was performed. They observed that the stainless-­steel reactor pro-
duces black powders, whereas the quartz reactor produced blur powders as shown
in Fig. 11.5 [31]. Zhang et al. used high concentration mixtures of H2/Ar for hydro-
genation of rutile TiO2. They used 1:4 ratio of H2:Ar mixture (20 to 80 sccm) at
350 °C for 1 h in a tubular furnace at ambient pressure [32]. Similarly, Zhu et al.
272 P. A. K. Reddy et al.

Fig. 11.5  Pictures of (a) native TiO2 nanotubes, H2/Ar gas-­treated TiO2 nanotubes in (b) stainless
steel container and c quartz reactor [31]. Reprinted with permission from [31]. Copyright (2012)
American Chemical Society

achieved hydrogenated anatase TiO2 nanotubes in a quartz tube under a high con-
centration mixture of continuous hydrogen (8 sccm) and argon (10 sccm) flux for
5 h at 400–600 °C [33].

Ambient Hydrogen–Nitrogen Treatment

Instead of using the ideal gas Ar, several researchers used inert gas N2 to prepare the
gas mixture for the hydrogenation treatment to prepare the black TiO2. The N2/H2
gas mixture beat the Ar/H2 mixture in terms of cost and ambient reaction conditions.
Similar to H2/Ar mixture, the H2/N2 mixture contains the H2 v/v% at about 5–10%.
He et al. performed low-­temperature hydrogenation of TiO2 films at 350 °C for 2 h
in a mixture gas of 6% H2 and 94% N2 [34]. Whereas, Wang et al. conducted hydro-
gen treatment on rutile TiO2 nanorods with anatase nanoparticles at slightly higher
concentration of hydrogen, i.e., 10% H2 and 90% N2 at 400 °C for 30 min [35]. The
reaction produced gray-­colored thin film contrary to the yellow-­colored thin films
observed by He et al. [34]. In another work, Zhu et al. synthesized black TiO2 using
the spillover effect of Pt nanoparticles. They produced hydrogenated Pt-­TiO2 by
treating the Pt-­loaded TiO2 (Pt impregnation followed by NaBH4 reduction) in an
8  v/v% H2/N2 mixture at 200–700  °C for 4  h reaction time (Fig.  11.6). In this
approach, the hydrogenation was achieved at reaction temperature as low as 160 °C
and the hydrogenation was dominant at higher temperatures due to the hydrogen
spillover from Pt to TiO2 particles [36].

Ambient Argon Treatment

In this approach, the black TiO2 is prepared directly by annealing the precursors in
an Ar atmosphere without the hydrogen gas. In the course of thermal treatment, the
hydrogen is generated in situ form the precursor or external additives and reduces
11  Black TiO2: An Emerging Photocatalyst and Its Applications 273

Fig. 11.6  Graphic image of two different spillover paths to synthesize black Pt–TiO2 nanoparti-
cles. Route 1: TiO2 hydrogenation followed by Pt loading. Route 2: H2 spillover and concurrent
reduction of Pt–TiO2 [36]. Reprinted with permission from [36]. Copyright (2016) The Royal
Society of Chemistry

the TiO2 to black TiO2. For an instance, Zhang et  al. used NaBH4 as an external
hydrogen source to convert the sol–gel-­derived Ni-­doped TiO2 to produce black Ni-­
doped TiO2 nanoparticles under Ar atmosphere at 350 °C for 1 h. The material’s
color changes from yellow to black when treated in Ar atmosphere and confirms the
hydrogenation of TiO2 [37]. In another approach, Grabstanowicz et al. used TiH2
and hydrogen peroxide as the starting materials. The yellowish initial powder
obtained by the TiH2 and H2O2 reaction at 100 °C was converted to black rutile TiO2
in an Ar atmosphere at 630 °C for 3 h [38]. Myung et al. prepared black TiO2 by
using the HF as the source of hydrogen. They prepared a yellow gel by the hydroly-
sis of TiCl4 in aqueous ethanol containing HF and urea. This yellowish powder was
annealed at 400–600 °C for 5 h in Ar gas to produce black TiO2 [39].

Argon–Nitrogen Treatment

This method produces black TiO2 without the usage of hydrogen. In this process, a
high-­temperature annealing produces oxygen vacancies in the lattice structure of
TiO2. Li et  al. developed black TiO2-­B nanoparticles by a tow step process [40].
First, the monodispersed TiO2-­B particles were obtained by the hydrolysis of TiCl4
in ethylene glycol and methanol solution under UV light irradiation in constant N2
bubbling. Then, the black TiO2-­B nanoparticles were obtained by annealing the fil-
tered solid product in Ar atmosphere at 340 °C for 2 h. In another report, Wei et al.
produced a core-­shell black TiO2 in a one pot preparation method by annealing the
colloidal TiO2 particles in N2 gas obtained by the partial hydrolysis of titanium
butoxide precursor [41]. The core-­shell structure of the as-­synthesized black TiO2
contains the surface oxygen vacancies with high concentration of Ti3+ as the shell
and the well-­ordered TiO2 lattice as the core (Fig. 11.7) [41].
274 P. A. K. Reddy et al.

Fig. 11.7  Synthesis mechanism of core-­shell black TiO2 form colloidal titania (obtained from
tetra butyl titanate and urea hydrolysis) annealed in N2 gas at atmospheric pressure [41]. Reprinted
with permission from [41]. Copyright (2017) Elsevier

11.2.2  Plasma Treatment

Although the synthesis of black TiO2 by hydrogenation at high temperatures is very


widely followed, there are substantial reports that high-­temperature hydrogenation
can be counterproductive for the overall performance of black titania [42]. Hence,
alternative strategies are developed to prepare black TiO2, and hydrogen plasma
treatment is one among them. There are several literature findings about black tita-
nia being synthesized by hot filament hydrogen plasma process. By applying 200 W
input power, Wang et  al. produced black TiO2 by hydrogen plasma in a furnace
operating at 500 °C for 4–8 h [43]. Alternatively, Teng et al. used hydrogen plasma-­­
assisted CVD (chemical vapor deposition) method to produce black TiO2. In this
method, the TiO2 powder was placed in corundum crucible under a hot filament
(2000 °C), and the reaction was performed in hydrogen plasma at 350–500 °C for
3 h [44]. Yan et al. used 3000 W inductively coupled plasma with 50 sccm H2 flow
rate and 26.5–28.3 mTorr to produce hydrogen plasma. The anatase black TiO2
nanoparticles obtained showed absorption in the visible light region [45]. In another
report, water plasma was used to oxidize the titanium metal electrodes (subjected to
high frequency bipolar voltage pulses). The resulting electrodes with surface HO−
and O2− functionalities were bombarded with the atomic hydrogen to produce the
black TiO2 [46].
11  Black TiO2: An Emerging Photocatalyst and Its Applications 275

11.2.3  Chemical Reduction

Metal Reduction

Black TiO2 can be prepared by the metal reduction method. This method is based on
the removal of oxygen atom from the TiO2 lattice structure at elevated temperatures
in presence of metal. In this method metal nanoparticles are used as the sacrificial
reductants wherein the metal undergoes oxidation by absorbing oxygen from the
TiO2 surface to produce black TiO2. Aluminum, zinc, magnesium, and lithium met-
als can be used as the reducing agents to produce black TiO2. Aluminum was used
by Wang et al. as the reductant in a two-­zone vacuum furnace at elevated tempera-
tures and low pressures less than 0.5 Pa [47]. The aluminum is heated at 800 °C in
one zone, and the TiO2 samples are placed at 300–600 °C on the other zone of the
furnace. Here the oxygen transfer takes place from the pre-­annealed TiO2 to the
molten aluminum creating oxygen vacancies in the TiO2 to produce black TiO2. In
another report, Cui et al. followed the same strategy in which the molten aluminum
was used to reduce the TiO2 nanotube arrays synthesized by the anodization method
over Ti foil [48]. The TiO2 nanotube arrays and powder aluminum were kept at two
different zones of the furnace and heated at 500 and 850 °C respectively. Lin et al.
further extended this strategy to produce nonmetal H, N, S, and I doped black TiO2
[49]. First, oxygen vacancies were created on the surface of TiO2 by aluminum
reduction method in which the TiO2 and aluminum powders were reacted at 500 °C
and 800 °C, respectively at 0.5 Pa pressure. The as-­obtained oxygen-­deficient black
TiO2 was hydrogenated at 500 °C for 4 h in hydrogen plasma to produce H-­TiO2.
Whereas, the S-­TiO2 and I-­TiO2 were prepared by reacting the black TiO2 with the
S and I2 at 500 °C for 4 h, respectively. N-­doped TiO2 was obtained by annealing the
black TiO2 in the 2:1 ratio NH3/Ar gas flow at 500 °C for 4 h. The produced samples
showed improved absorption in the visible and infrared regions.
Zhao et al. proposed a solvothermal method to produce black TiO2 in which zinc
metal powders were used as the reducing agents [50]. In this method, TiCl3 aqueous
solution was reacted with Zn powders in isopropanol solution at 180 °C for 6 h in
an autoclave. Here the Zn powders supposed to prevent the complete oxidation of
the Ti3+ ions during the reaction. They also observed that the color of the final prod-
uct can be manipulated form gray to dark blue by simply changing the amount of
Zn. Sinhamahapatra et al. developed a method to produce black TiO2 by using mag-
nesium as the reducing agent [51]. In this method, the commercial TiO2 and magne-
sium powders were well-­mixed and placed in a tubular furnace at 650  °C under
5 v/v% H2/Ar gas flow for 5 h. After the reaction, the excess magnesium metal was
removed by reacting with 1.0 M HCl followed by washing with excess water.
Zhang et  al. reported a novel method to selectively produce rutile black TiO2
using Li as the reductant in an ethanediamine solution [52]. They dissolved lithium
foils in ethanediamine solution and soaked the TiO2 for 6 h under continuous stir-
ring. The product was washed with 1 M HCl and water to produce black rutile TiO2.
However, this method cannot be applied to reduce the anatase TiO2.
276 P. A. K. Reddy et al.

Reduction by Hydride Slats

Black TiO2 can be synthesized by chemical reduction method using strong reducing
agents such as NaBH4. A room temperature reduction of TiO2 nanotubes prepared
from anodization of Ti foil was carried out by Kang et al. [53]. The experiment was
performed in a 0.1 M NaBH4 solution for 10–60 min. Whereas, Tan et al. developed
a solid-­state reaction method in which the pristine TiO2 was ground with the NaBH4
reductant and was subjected to annealing in Ar gas at 300–400 °C up to 1 h [54]. In
another work, CaH2 was also used as the reducing agent by Tominaka et al. to syn-
thesize black TiO2. The preparation was carried out at 350 °C temperature in which
the rutile TiO2 nanoparticles were annealed with CaH2 powder for 10–15 days to
obtain black rutile TiO2 nanoparticles [55]. Zhu et al. prepared crystalline-­­amorphous
core-­shell black TiO2 by reducing Degussa P25 TiO2 by CaH2 at 400 °C. The syn-
thesized black TiO2 showed abundant oxygen vacancies that led to improved solar
absorption and photocatalytic activity [56].

11.2.4  Electrochemical Synthesis

Black TiO2 nanotube arrays can be prepared by electrochemical anodization method


in aqueous media. Zhang et al. developed a two-­step anodization followed by elec-
trochemical reduction method to produce black TiO2 nanotube arrays on Ti sheet
[56]. In the first anodization step, the TiO2 nanotube arrays were obtained by anod-
izing the Ti sheet at 60 V for 30 min in ethylene glycol solution containing 0.5 w%
NH4F and 2 v/v% H2O2 using Pt as counter electrode. The as-­produced TiO2 nano-
tubes were removed by ultrasonication, and then the second anodization step was
performed at 80  V for 5  min in the similar solution. The TiO2 nanotubes were
annealed at 450 °C for 1 h. Then, the electrochemical reduction was conducted in
1 M Na2SO4 solution for 30 min by applying the negative potential (0.4 V vs RHE)
[57]. Zheng et al. synthesized the black TiO2 films by using a multipulse anodiza-
tion method [58]. Dong et  al. prepared black TiO2 by anodization followed by
annealing as portrayed in the Fig. 11.8 [59]. The TiO2 nanotubes were grown on Ti
foil by anodizing at 60 V in ethylene glycol containing 0.25 wt% NH4F and 2 v/v%
H2O for 10  h. The nanotubes were removed, and a second anodization step was
performed with similar conditions. Finally, the anodized Ti foil was washed with
ethanol and water and dried and subjected to annealing at 450 °C for 1 h. The as
produced substrate contains a layer of black TiO2 as shown in the Fig. 11.8 [59].
11  Black TiO2: An Emerging Photocatalyst and Its Applications 277

Fig. 11.8  Experimental process and optical images of the stripped TiO2 layer [59]. Reprinted with
permission from [59]. Copyright (2013) The Royal Society of Chemistry

11.2.5  Pulsed Laser Ablation

Black TiO2 preparation by pulsed laser ablation method is based on the creation of
defect sites on the surface of TiO2 by applying high intensity laser beam pulses.
Chen et al. applied Nd:YAG laser pulses on to the TiO2 suspended in the aqueous
solution to synthesize black TiO2 [60]. After applying the laser pulsed for 120 min,
they found the white pristine TiO2 turn into black TiO2. The laser irradiation removes
the oxygen creating surface oxygen vacancies that leads to the formation of black
TiO2−x [61]. Lü et al. obtained a bilayer amorphous black TiO2/crystalline TiO2 film
by applying laser pulses on to a commercial TiO2 target at 100 °C for 10 min in
vacuum by using KrF excimer laser at a repetition rate of 2 Hz with a laser fluence
of 2 J.Cm−2. A metallic conduction between the amorphous and crystalline interface
via electronic interface reconstruction [62].

11.2.6  Other Methods


Ionothermal Synthesis

Li et al. developed a method to synthesize single crystalline black TiO2 using the
ionic liquid. In a typical experiment, the Ti metal foil was treated in a buffer solution
of lithium acetate, and acetic acid dissolved in N-­N-­dimethylformamide containing
1-­methyl-­imidazolium tetrafluoroborate (ionic liquid) for 200  °C for 24  h [63].
They observed that the ionic liquid with fluorine and acetic acid involve in the
278 P. A. K. Reddy et al.

dissolution of Ti foil followed by the oxidation, respectively. The in situ generated


hydrogen (by the reaction of Ti foil with the acetic acid) hydrogenates the TiO2
surface to produce H-­TiO2−x disordered layer. This layer composed of Ti3+ enriched
black TiO2 showed exceptional optical properties under visible light.

Imidazole Reduction

Zou et al. used imidazole as the reductant to produce gray TiO2 [64]. They intro-
duced the mixture of amorphous TiO2, imidazole, and HCl into a preheated muffle
furnace operating at 450 °C and anneal the mixture for 6 h. The obtained gray col-
ored reduced TiO2 showed strong absorption through the entire visible spectrum.

Proton Implantation

Proton implantation method was implemented by Liu et al. to produce black TiO2
[65]. They carried out the experiment at a 30 keV energy and 1016 ions.cm−2 dosage
using Varian 350 D ion implanter. The ion implantation selectively modified the
TiO2 nanotubes at their tops and induced defects. The proton implanted top surface
with defects acts as intrinsic co-­catalytic centers. They propose synergism between
the implanted upper part (acts as a catalyst) and the intact implant-­free lower part
(acts as a light absorber) for enhanced activity.

Si Quantum Dot (QD)-Assisted Chemical Etching

Huang et al. reported the Si chemical etching method to introduce defects contain-
ing mesoporous TiO2 on the Ti foil. The hydrogen terminated Si QDs were depos-
ited on the surface of the Ti foil by electrodeposition technique. Later, the Si QDs
were etched with the HF wherein the hydrogen terminated Si QDs facilitate the
formation of Ti3+ alongside with the oxygen defects. Ultimately, a mesoporous
black TiO2 surface was formed on the surface of Ti foil [66].

11.3  Black TiO2 Properties

Titanium dioxide is a semiconductor metal oxide that exists in four different phases,
namely, anatase, rutile, brookite, and TiO2-­B. All these polymorphs have same TiO6
octahedra however differ in the distortion of octahedron units and share edges and
corners of the octahedron units in different ways. So far various black TiO2 nanoma-
terials have been synthesized from white pristine anatase, rutile, brookite, and
TiO2-­B. In addition, different preparation conditions and fabrication methods were
adopted to produce black TiO2; it is also reported that these various black TiO2
11  Black TiO2: An Emerging Photocatalyst and Its Applications 279

showed different physical/chemical properties. A detailed knowledge of these prop-


erties is essential to understand the nature of the black TiO2 nanomaterials in terms
of their chemical, optical, and catalytic behaviors.

11.3.1  Existence of Surface Disorders

The color change from white to black (during the black TiO2) reflects the alteration
in the optical properties caused by the structural disorders at the surface of the black
TiO2 nanomaterials prepared by different methods. Some examples are the crystal-
line/disordered core-­shell structure observed for black TiO2 prepared by hydrogena-
tion [24, 29, 67, 68], hydrogen plasma [43], and aluminum reduction [47] methods.
Many reports attempted to investigate these disorders in the crystalline phase struc-
ture of black TiO2 by comparing the XRD patterns [23, 24, 29]. It is obvious that the
XRD peaks in black TiO2 tends to show shift due to the structural disorders. In the
diffraction patterns, black TiO2 showed peak shift to higher angles suggesting a
reduced interplanar distance in the crystalline phase [67, 68]. Since XRD is a bulk
technique that is sensitive to crystallinity, several researchers tried Raman spectros-
copy to identify the disordered structures or amorphous nature of the black TiO2
surface [23, 68]. Many studies found that the black TiO2 nanomaterials with disor-
ders or amorphous nature tends to show weak intensity [33, 69, 70], broad width
peaks [24, 29, 39, 47] with a shift to higher wavenumber [42, 48, 71], and additional
vibrational modes [23, 64, 69] in the Raman spectra as compared to the pristine
white TiO2.
High-­resolution transmission electron microscopy (HRTEM) was employed to
visualize the structural disorders in the black TiO2 surface. For an instance, Chen
et al. compared the HRTEM and line analysis of the white TiO2 and black TiO2 [72].
The white TiO2 nanoparticles showed well-­resolved, perfect lattice fringes with uni-
form interplanar distance (0.352 nm for anatase) all over the crystallite including
bulk and surface (Fig. 11.9a, b). Whereas, the black TiO2 had a disordered core-­shell
structure at the outer layer where the interplanar distance is not uniform (Fig. 11.9c,
d) [72]. The selected area electron diffraction (SAED) further revealed the crystal-
line/amorphous core-­shell structure of the black TiO2. The SAED patterns of the
black TiO2 showed cloudy diffraction patterns contrary to the clear diffraction pat-
terns (made of diffraction dots/rings) observed for the pristine white TiO2 [73].
Attempts were also made to evaluate the exact volume and percentage of these dis-
orders in the black TiO2 by combining the information gained from the XRD,
Raman, and HRTEM.
280 P. A. K. Reddy et al.

Fig. 11.9 (a, b) (HRTEM) and (c, d) (line analyses) of white and black TiO2 nanocrystallites,
respectively. The zeros of the axis in (b, d) correspond to the left ends of the lines in (a, c). The red
and green curves in (b, d) correspond to the red and green lines in (a, c) [72]. Reprinted with per-
mission from [72]. Copyright (2013) Nature Publishing Group

11.3.2  Existence of Ti3+ Centers and Oxygen Vacancies

Presence of defect sites such as Ti3+ and oxygen vacancies (Vo) in a material can be
identified by characterization techniques such as X-­ray photoelectron spectroscopy
(XPS), electron spin resonance spectroscopy (ESR spectroscopy), and synchrotron
X-­ray absorption–emission photoelectron spectroscopy. The Ti3+ centers can be dis-
tinguished by XPS; however, their existence in black TiO2 is solely dependent on
the synthesis procedure. For example, Ti3+ ions were not detected for the black TiO2
nanomaterials (prepared by hydrogen reduction and hydrogen plasma treatment)
with the abovementioned techniques [72, 74, 75]. However, several other reports
claimed the existence of Ti3+ ions in black TiO2 prepared by methods such as hydro-
gen treatment [27, 29, 76], chemical reduction [48, 77], chemical oxidation [38],
and electrochemical reduction [57, 59].
Oxygen vacancies are the other structural defects frequently reported in the black
TiO2 prepared by the hydrogen thermal treatment [69, 78, 79], electrochemical
reduction [38, 80], chemical reduction [64, 77, 81], and chemical oxidation [82].
For instance, ESR spectroscopy revealed oxygen vacancies in the black TiO2
nanoparticles prepared by Al metal reduction method [77, 81], thermal hydrogen
treatment [22, 83], and electrochemical reduction [59]. However, oxygen vacancies
11  Black TiO2: An Emerging Photocatalyst and Its Applications 281

were not observed for black TiO2 in some cases. Xia et al. prepared black TiO2 with
thermal treatment did not observe the oxygen vacancies in the ESR spectra [75].

11.3.3  Existence of Surface Functional Groups

During the synthesis of black TiO2, its surface of undergoes numerous changes
especially the alterations in the surface functional groups. These functional groups
play important role in influencing the surface chemistry of TiO2 that ultimately
affects the activity. Black TiO2 nanomaterials prepared by hydrogenation treatment
showed increased concentrations of OH groups. This OH groups are characterized
by a shoulder peak in the O 1 s XPS spectrum related to the Ti-­OH bonding. Black
TiO2 nanoparticles prepared by treating TiO2 nanoparticles with ultrahigh pure
hydrogen in the temperature range of 200–600 °C showed Ti-­OH surface functional
groups [23, 28, 74, 76]. Conversely, Xia et al. observed a decrease in the content of
surface OH groups in the O 1 s XPS spectrum of black TiO2 prepared in 5 v/v% H2/
Ar atmosphere at 450  °C for 4  h at 5  bar pressure [73, 75]. The OH functional
groups are also characterized by the Fourier transform infrared (FTIR) spectrum of
black TiO2 nanomaterials by the variations in the OH vibrational band [43, 70, 73,
75]. Black TiO2 prepared by hydrogen plasma treatment at 500 °C for 4–8 h dis-
played various peaks from 3600 to 3800 cm−1 related OH functionalities with differ-
ent chemical environments. In contrast, the reduced OH peak intensity was shown
for black TiO2 prepared in 5 v/v% H2/N2 gas mixture at 500 °C for 1 h and 5 v/v%
H2/Ar atmosphere at 450 °C for 1 h [70, 84]. Similar observations were made by
Chen et al. and Xia et al. for black TiO2 produced by hydrogenation method [72, 73,
75]. 1H nuclear magnetic resonance (NMR) spectra were employed by several
researchers to identify the OH groups in black TiO2. Wang et al. observed peak at
5.5 ppm by the bridging hydroxyl groups and two more peaks at 0.01 and 0.4 ppm
related to the internal and terminal hydroxyl groups for the black TiO2 prepared by
hydrogen plasma treatment at 500  °C for 4–8  h [43]. Similarly, a broad peak at
5.7 ppm and two additional peaks at −0.03 and 0.73 ppm were observed by Chen
et al. for the hydrogenated TiO2 [72].
The existence of titanium bonding with hydrogen in black TiO2 was also reported
by several reports [43, 68, 74, 84]. A shoulder peak in the Ti 2p spectrum at 457.3 eV
due to the surface Ti-­H bonds were observed by Zheng et al. and Wang et al. for the
black TiO2 prepared by hydrogenation method at 500 °C in 5 v/v% H2/N2 atmo-
sphere and hydrogen plasma treatment at 500 °C for 4–8 h, respectively [43, 84].
Wang et al. suggested the Ti-­H bonds cover the black TiO2 surface prepared by the
hydrogenation of TiO2 in ultra-­pure H2 atmosphere at 400 °C for 2 h [68]. The Ti-­H
bond was also substantiated by the peak at 59.3° in the XRD pattern of black TiO2
prepared in ultrapure H2 [74].
282 P. A. K. Reddy et al.

11.3.4  M
 odifications in the Band Structure, Color,
and Electron Trap Sites

The general observation during the synthesis of black TiO2 is that the color transfor-
mation of pristine white TiO2 occurred. The extent of color change depends on the
factors like extent of hydroxylation and hydrogenation, doping content, and synthe-
sis conditions such as volume of reducing agents (H2/Ar/N2, NaBH4, Al, Zn etc.)
used, thermal reduction temperature, reaction time, etc. During the hydrogenation
approaches, the presence of H atoms on the surface and inside the subsurface layer
promoted both H and H2 penetration into the subsurface layer and prevented the
escape of the H2 from the cage. The H2 molecule inside a cage readily dissociated
and formed 2HO-­species exothermically, which further transformed into H2O and
resulted in the formation of O-­vacancies and surface disordering. Further insights
into the shell structure have yielded some interesting findings. It is to be observed
that the outmost layer of black TiO2 nanoparticles consists of a disordered Ti2O3
shell. It is believed that there is a transition region that connects the disordered Ti2O3
shell to the perfect rutile core consisting first of four to five monolayers of defective
rutile contain visible Ti interstitial atoms, which is followed by an ordered recon-
structed layer of the Ti interstitial atoms [85]. The interstitial H, H2, and H2O brought
in density of states within the TiO2 bandgap and induced a shift of the bandgap posi-
tion notably toward the conduction band. The atoms not only induced the lattice
disorders but also interacted strongly with the O 2p and Ti 3d states, resulting in a
considerable contribution to the mid-­gap states (Fig. 11.10). The optical absorption
was dramatically red shifted due to the mid-­gap states and the photogenerated elec-
tron hole separation. In addition to hydrogenation-­based approaches, the other
approaches have also been tested such as reduction and other and partial oxidation

Fig. 11.10  Black titania surface structure


11  Black TiO2: An Emerging Photocatalyst and Its Applications 283

approaches. For instance, Zhang et al. have prepared a disordered rutile black titania
by employing a reducing agent, lithium in ethylenediamine (Li-­EDA) [52].
Even in the black titania oxidation-­based preparation approaches, some reduc-
tant are added to avoid complete oxidation of the low valence Ti species. Black
titania exhibits unique characteristics based on type of preparation approaches. For
example, Ti-­H bonds on the black titania surface are predominantly observed due to
hydrogen reduction-­based approach. Furthermore, the mode of the surface oxygen
vacancies formation varies based on whether it’s an oxidation or reduction strate-
gies. In the case of the reduction process, the oxygen atoms on the surface of TiO2
can be removed by reductant resulting in the formation of the surface oxygen vacan-
cies. Whereas in the partial oxidation-­based approach, the solutions that are used in
oxidative preparation of the black titania could lead to oxygen vacancy/Ti3+ centers
in the bulk TiO2 matrix (Fig. 11.10). The H-­doping and oxygen vacancy/Ti3+ usually
incorporated in TiO2 results in mid-­gap states which renders the colorization of the
TiO2 there by altering the optical properties. Further the mid-­gap states also act as
trapping sites for the electrons and holes which avoids the recombination process.
The surface Ti4+-­OH groups act as trapping centers for the electrons and the O2−
resulted from the adsorbed oxygen on the oxygen defective site acts as a hole trap-
ping center (Fig. 11.10).
Thus, as discussed above, the induction of important characteristics like defec-
tive and disordered core-­shell structure, oxygen vacancy, and surface functional
groups such as OH play important role in changing the electronic and optical prop-
erties of black titania [29]. Bandgap narrowing is a direct result of the changes in the
band structure of black TiO2. The origin of bandgap reduction is debated and dis-
cussed for a long time due to the complexity of factors such as defects, disorders,
doping, and surface functional groups in black TiO2. Some assumptions believed
that the Ti3+ species might be responsible for the bandgap reduction [38, 86–88] and
mid-­gap states formed by the oxygen vacancies promote the low energy light
absorption (below direct bandgap) by indirect electronic transitions [86]. Conversely,
the other assumptions supposed that the bandgap narrowing is due to the disorder/
defect induced mid-­gap states that significantly upshift the valance band (VB) edge
and that the conduction band (CB) tails states arising from the surface disorders
could marginally contribute to the bandgap narrowing of black TiO2. Chen et  al.
thought that the bandgap narrowing mainly originated from the disorder-­induced
mid-­gap states rather than the Ti3+ species, greatly upshifting the valence band (VB)
edge of black TiO2, and that the possible conduction band (CB) tail states arising
from the surface disorder could only slightly narrow the bandgap (Fig. 11.11a) [23,
72]. Naldoni et al. supplemented this claim by stating that the localized states are
created at 0.7–1.0 eV below the CB minimum due to the oxygen vacancies in the of
black TiO2 [29]. Figure 11.11b illustrates comparison of the density of states in the
black and pristine white TiO2. Wang et al. further clarified that both the CB and VB
tails responsible for slight reduction in the bandgap by 0.8  eV and the mid-­gap
states are induced by the Ti-­H bonding at 0.92–1.37 eV below the CB minimum of
balck TiO2 [43]. To conclude, the structural changes in the bandgap of black TiO2
284 P. A. K. Reddy et al.

Fig. 11.11  Illustration of density of states (DOS) of a black and pristine white TiO2, b oxygen
deficient TiO2 (TiO2−x) and hydrogenated black TiO2 (TiO2−xHx)

primarily endorsed the CB and VB tails along with the induced mid-­gap states by
the oxygen vacancies or Ti-­H bonds [23, 29, 43, 44].

11.4  Applications of Black Titania in Photocatalysis

Black titania has wide range of applications in the field of photocatalysis. Due to its
higher activity, they have been tested in major energy and environmental applica-
tions which are explained.

11.4.1  Pollutant Removal

The unique properties of the black titania has helped in the effective degradation of
the range of aqueous pollutants through the photocatalysis process. For instance, in
a work carried out by the Teng et al., the black TiO2 was able to successfully degrade
organic molecules in water under solar illumination [44]. This enhanced perfor-
mance was attributed to the oxygen vacancies and Ti-­H on the surface. In a work
carried out by Zhang et al., the uniform black TiO2 nanothorns/graphene/black TiO2
nanothorns sandwich-­like structured nanosheets are prepared [89]. The black TiO2
nanothorns that are present on the either sides of the graphene are prepared; this
sandwich-­like structure-­based photocatalyst was able to degrade the 99% of the
pesticide atrazine. The overall better performance was attributed to the surface dis-
orders of the black TiO2, as well as it is believed that sandwich like structure would
probably has the role in separation and transportation of photogenerated charge
carriers, thus increasing the efficiency. In another work, the novel Ni2+ and Ti3+ co-­­
doped porous black anatase TiO2 was prepared. It was observed that Ni2+/Ti3+ co-­­
doped black TiO2 had narrow bandgap due to the formation of mid-­gap states which
helped the catalyst to utilize visible light as well as prevented recombination. The
overall degradation ratio of methyl orange and rhodamine B is up to 95.38 and
95.86%, resp., within 150 min irradiation of visible light [37]. As it is well known
11  Black TiO2: An Emerging Photocatalyst and Its Applications 285

that hydrogenated-­based reduction of the TiO2 was a prime avenue for the prepara-
tion of the conversion of the pristine TiO2 into the black TiO2, it was proven that
slight hydrogenation was more beneficial to the improving of the performance of
the black TiO2. This was experimentally proven by the degradation studies wherein
the methylene blue dye was more efficiently degraded by the slightly hydrogenated
TiO2 in comparison to the pristine as well higher hydrogenation TiO2. This poor
performance of the higher hydrogenated TiO2 was due to higher bulk defects that in
turn resulted in the decreased number of trapped holes and higher recombination. In
a work, the molybdenum disulfide was sandwiched between mesoporous black TiO2
with either sides; thus the heterojunctions are formed between MoS2 and mesopo-
rous black titania [90]. This structure was able to degrade 89.86% of methyl orange
dye. The enhanced performance can be attributed to decreased bandgap separation
and transport of the photogenerated charges [90].

11.4.2  Hydrogen Production

The photocatalysis-­based hydrogen production is one of the important applications


of the renewable energy research. In the recent years, the titania-­based modifica-
tions have substantially improved the feasibility of our chances in the water splitting
work. Black titania-­based water splitting work also had attracted huge attention.
Hydrogen plasma-­based approach was been employed to synthesize unique H-­doped
black titania with a core-­shell structure (TiO2@TiO2−x). As a resultant, the plasma-­­
based approach yielded superior performance in comparison to the high H2-­pressure
process. H doping is favorable to eliminate the recombination centers of light-­­
induced electrons and holes [87]. In addition to the conventional black titania, vari-
ous modifications have been made to test their overall improvement in the efficiency
of the hydrogen evolution through the water splitting. It must be noted that prior to
hydrogenation, the digestion of the titania under strong basic conditions such as
NaOH has led to higher hydrogen capacity [91]. In the recent years, the mesoporous-­­
based titania has been used as parent material which is further hydrogenated to yield
black titania material was actively pursued. In a work carried out by Hu et al. [92],
the stable mesoporous black titania hollow spheres were synthesized by high tem-
perature hydrogenation combined with amine encircling process. According to the
experimental results the hydrogen evolution was (241 μmol.h−1) which was twice
the efficiency in comparison to bare black titania. Similar results about mesoporous
titania were reported in another experimental studies, wherein the hydrogen evolu-
tion rate of the mesoporous black titania yielded (136.2 μmol.h−1), which is twice as
high as that of pristine mesoporous TiO2 (76.6 μmol.h−1) [93]. In another work, the
magnesiothermic reduction has been developed to synthesize reduced black TiO2
under a 5% H2/Ar atmosphere. This developed material was able to yield good
hydrogen evolution performance in the methanol–water system in the presence of Pt
as a co-­catalyst. The maximum hydrogen production rates are 43 mmol.h−1.g−1 and
440 μmol.h−1.g−1 under solar illumination [51].
286 P. A. K. Reddy et al.

11.4.3  Photo Reduction of CO2

The photocatalysis-­based reduction of carbon dioxide into fuels with H2O, under
the solar illumination, has been attracting a lot of attention in the recent years. The
utilization of the black titania in its brookite form was subjected to oxidation-­based
hydrothermal treatment combined with post annealing. It was observed that the
brookite titania prepared by this method was able to induce photoreduction, and it
yielded 11.9  μmol.gcat−1.h−1 for CH4 and 23.5  μmol.gcat−1.h−1 for CO [94]. In
another work, Black TiO2 films with unique porous structures were used as photo-
catalysts to reduce CO2 under simulated sunlight irradiation at room temperature in
the presence of H2O. These films have shown extremely high CO2 photoreduction
efficiency with favorable selective formation of CO and CH4 (with highest produc-
tion rates of 115 and 12 μmol.g−1.h−1, respectively) compared to conventional TiO2
(Degussa P25) (0.28 and 0.019 μmol.g−1.h−1, respectively) [95]. Even the coating of
metal nanoparticles can potentially improve the performance of the black titania-­­
based photocatalysts. For instance, in a work carried out by Zhao et al., black TiO2-­­
coating process Cu nanoparticles (denoted as Cu@TiO2) in the photoreaction of
CO2 with H2O vapor under visible light irradiation. The photocatalytic activity for
Cu@TiO2 (∼4%Cu) reaches 1.7 times of that for its counterpart, bared black TiO2
[96]. In addition to the photo reduction, efficient visible light photocatalytic CRM
by combining Pt/black TiO2 catalyst with light-­diffuse-­reflection-­surface. Under
visible light illumination by filtering UV light from AM 1.5G sunlight, H2 and CO
yields reached 71 and 158 mmol.h−1.gcat−1, with a quantum efficiency of 32.3% at
550 °C, and 129 and 370 mmol.h−1.gcat−1, with a quantum efficiency of 57.8% at
650  °C.  Those yields are three orders of magnitude larger than the reported val-
ues [97].
Table 11.1 reviews the several black TiO2 nanomaterials with a different prepara-
tion method, their properties, and photocatalytic applications. As it is seen from the
table, the electronic, optical, structural, and chemical properties black TiO2 exten-
sively depend on the synthesis methods; it is very important to pay attention to the
synthesis procedures to achieve the desired properties for specific applications.

11.5  Conclusions and Future Outlook

Black titania is an excellent photocatalyst with enhanced performance due to its


capacity to absorb higher visible irradiation, oxygen vacancies, mid-­gap state for-
mation, valence band movement, and so on. Based on available literature, it is
observed that hydrogenation-­based treatment of the black titania preparation is
more prevalent. Although the hydrogenation-­based approach is proven successful,
considering the disadvantages of high temperatures in the hydrogenation the alter-
native approaches are encouraged which can attain better efficiency in the perfor-
mance of the black titania photocatalyst. Based on the literature findings, it is clearly
11  Black TiO2: An Emerging Photocatalyst and Its Applications 287

Table 11.1  Performance evaluation of the recently reported black titania-­based photocatalysts
Sl Crystalline Active Photocatalytic
no Synthesis method phase Morphology centers application Ref
1 High pressure H2 Anatase Core-­shell Ti3+, Vo Methylene blue (MB) [98]
and phenol
degradation
2 High pressure H2 Anatase Vo MB degradation [42]
3 High pressure H2 at Mixed phases Ti3+ H2 production [24]
room temperature of anatase,
rutile, and
Ti2O3
4 Low pressure H2 Rutile Nanowire Vo Photoelectrochemical [28]
arrays water splitting
5 Hydrogen by using Anatase Mesoporous, Ti3+ H2 production [99]
bio-­template nanosheets
6 H2-­Ar Anatase Nanotube Ti3+, Vo H2 generation [22]
arrays
7 H2-­Ar MB degradation [100]
10 Magnesiothermal Anatase Ti3+, Vo H2 production [51]
H2-­Ar
11 UV irradiation of Anatase and Ti3+, Vo Methyl orange (MO) [40]
solution followed TiO2-­B degradation
by Ar treatment
12 N2 gas at Anatase Core-­shell Ti3+, Vo MO degradation [41]
atmospheric
pressure
13 H2/N2 atmosphere Mixed phase Core-­shell Vo Water splitting [36]
of anatase and
rutile
14 Low pressure H2 or Anatase and Ti3+, Vo MO degradation [101]
H2/N2 rutile
15 H2/N2 flow Temperature Core-­shell Ti3+, Vo Phenol degradation [102]
dependent; nanotubes
anatase and nanobelts
(<150 °C) and
TiO2-­B
(>150 °C)
16 Ni2+ doping Anatase Core-­shell Ti3+, Vo MO and Rhodamine [37]
followed by Ar (RhB) degradation
treatment
17 TiH2 oxidation by Rutile Ti3+, Vo MB degradation [38]
H2O2 then Ar
treatment
18 Hydrogen plasma Anatase Core-­shell Ti3+, Vo H2 generation [43]
19 H2 plasma CVD Mixture of Core-­shell Vo RhB degradation [44]
Na0.23TiO2 and
Ti2O3
20 H2 plasma Anatase Core-­shell Ti3+, Vo MB degradation [103]
(continued)
288 P. A. K. Reddy et al.

Table 11.1 (continued)
Sl Crystalline Active Photocatalytic
no Synthesis method phase Morphology centers application Ref
21 H2 plasma Anatase and Nanoporous Vo, Ti3+ Phenol, reactive black [104]
brookite 5 (RB 5), Rho B and
mixed phase MB degradation
22 Water plasma Anatase and Nano-­/ Ti2+, MB degradation [46]
rutile mixed microspheres Ti3+, Vo
phase
23 NaBH4 reduction Ti3+, Vo H2 production [53]
24 NaBH4 reduction Anatase and Core-­shell Ti3+, Vo H2 generation [54]
rutile mixed
phase
25 NaBH4 reduction + Mixture of Nanobelts Ti3+, Vo MO degradation and [105]
Ar treatment Ti2O3, Ti3O5 H2 generation
and TiO2
26 NaBH4 + N2 Anatase Ti3+, Vo H2 production [106]
annealing
27 Al metal reduction Anatase Core-­shell Ti3+, Vo MO degradation and [47]
H2 generation
28 Al metal reduction Anatase Nanotube Ti3+, Vo Photoelectrochemical [48]
arrays water splitting
29 Al reduction Anatase and Core-­shell Ti3+, Vo MO degradation and [49]
rutile mixed H2 generation
phase
30 Mg metal reduction Anatase and Core-­shell Solar water [107]
rutile mixed evaporation
phase
31 Ionothermal Anatase Single crystal Ti3+, Vo RhB and aniline [63]
method degradation
32 Electrochemical Anatase Nanotube Vo Photoelectrochemical [108]
reduction water splitting
33 Electrochemical Mixture of Nanotube Ti3+ Water splitting [57]
reduction anatase, Ti4O7 arrays
and Ti6O11
34 Electrochemical Anatase Nanotubes Ti3+, Vo RhB degradation [109]
doping
35 Electrochemical Anatase Hexagonally Ti3+, Vo RhB degradation [59]
anodization dimpled layer
36 Pulsed laser Anatase Nanospheres Ti3+, Vo RhB degradation [60]
ablation
37 UV laser pulse Anatase and Ti3+, Vo Solar water splitting [61]
irradiation rutile mixed
phase
38 Si QD chemical Anatase Mesoporous Ti3+, Vo MO degradation [66]
etching film
39 Microwave Core-­shell Ti3+ RhB degradation [110]
(continued)
11  Black TiO2: An Emerging Photocatalyst and Its Applications 289

Table 11.1 (continued)
Sl Crystalline Active Photocatalytic
no Synthesis method phase Morphology centers application Ref
40 Hydrothermal/ Anatase and 3D flowers Ti3+ MO degradation and [111]
chemical reduction rutile mixed H2 generation
phase
41 Ultrasonic Anatase Core-­shell Acid fuchsin (AF) [112]
irradiation degradation
42 H2 plasma TiO2-­B Nanowires Vo Photoelectrochemical [113]
reduction water splitting
43 H2 plasma Rutile Nanowire Vo MB degradation [114]
arrays
44 Black TiO2 and Anatase Nanothorns Ti3+ Atrazine [89]
graphene decomposition
sandwich-­like
structures
45 Mg reduction Anatase, Core-­shell Ti3+ Photoelectrochemical [115]
rutile, and water splitting
mixed phases
of both
46 Al reduction/ Anatase and Core-­shell Ti3+, Vo MO degradation [116]
annealing rutile mixed
phase
47 Solvothermal Anatase Ti3+, Vo MB degradation [117]
reaction/annealing
48 Solvothermal Zn Anatase and Truncated Ti3+, Vo H2 production [50]
metal assisted Rutile mixed octahedron
reduction phase
49 H2 reduction Anatase Inverse opals, MB degradation [118]
Core-­shell
50 H2 plasma Anatase Nanofibers Ti3+, Vo Photoelectrochemical [17]
evaluation
51 Hydrogenated Anatase Microsphere H2 generation [84]
titanate nanotubes
52 Al metal reduction Brookite Core-­shell, Ti3+, Vo MB and MO [81]
flowers like degradation
53 Hydrothermal Anatase and Porous thin Ti3+, Vo CO2 reduction [95]
treatment rutile mixed films
phase
54 Hydrogenation Anatase Ordered Ti3+ H2 generation [93]
mesoporous
arrays
55 Solvothermal Anatase Mesoporous Ti3+ H2 production [92]
method hollow
spheres
56 Reduction in H2 at Anatase and Core-­shell Ti3+, Vo Atrazine degradation [119]
atmospheric rutile mixed
pressure phase
(continued)
290 P. A. K. Reddy et al.

Table 11.1 (continued)
Sl Crystalline Active Photocatalytic
no Synthesis method phase Morphology centers application Ref
57 Al reduction/H2S Anatase and Core-­shell Ti3+, Vo MO degradation and [120]
treatment rutile H2 generation
58 Proton Anatase Nanotube Ti3+ H2 production [65]
implantation arrays
59 Cu/H2 gas flow Anatase Vo CO2 reduction [96]
treatment
60 Hydrothermal/ Brookite Single crystal Ti3+, Vo CO2 reduction [94]
annealing in N2
flow
61 Anhydrous Anatase, Vo Disinfection [121]
lithium-­ethylene rutile, and
diamine (Li-­EDA) mixed phases
reduction in N2
atmosphere at
room

evident that black titania has surely better performance than TiO2. It was also
observed that when porous titania materials are converted to black titania based, the
performance enhances further. Furthermore, the black titania synthesized can be
further modified by using other approaches such as metal doping, composites, and
dye sensitization to test if it can further enhance the performance. The available
findings in this black titania research are still at inchoate stage; we call upon more
research in the utilization of the black titania in various photocatalysis-­based appli-
cations such as pollution removal, water splitting, and various other applications.

Acknowledgments  This study was supported by the 2017 Research Fund (1.170013.01) of
UNIST (Ulsan National Institute of Science & Technology).

References

1. Park, K., Zhang, Q., Myers, D., & Cao, G. (2013). Charge transport properties in TiO2 net-
work with different particle sizes for dye sensitized solar cells. ACS Applied Materials &
Interfaces, 5(3), 1044–1052.
2. Wu, S., Weng, Z., Liu, X., Yeung, K. W. K., & Chu, P. (2014). Functionalized TiO2 based nano-
materials for biomedical applications. Advanced Functional Materials, 24(35), 5464–5481.
3. Reddy, P. A. K., Reddy, P. V. L., Kim, K. H., Kumar, M. K., Manvitha, C., & Shim, J. J. (2017).
Novel approach for the synthesis of nitrogen-doped titania with variable phase composi-
tion and enhanced production of hydrogen under solar irradiation. Journal of Industrial and
Engineering Chemistry, 53(25), 253–260.
4. Reddy, J. K., Lalitha, K., Reddy, P. V. L., Sadanandam, G., Subrahmanyam, M., & Kumari,
V. D. (2014). Fe/TiO2: A visible light active photocatalyst for the continuous production of
hydrogen from water splitting under solar irradiation. Catalysis Letters, 144(2), 340–346.
11  Black TiO2: An Emerging Photocatalyst and Its Applications 291

5. Banerjee, B., Amoli, V., Maurya, A., Sinha, A. K., & Bhaumik, A. (2015). Green synthesis of
Pt-doped TiO2 nanocrystals with exposed (001) facets and mesoscopic void space for photo-­
splitting of water under solar irradiation. Nanoscale, 7(23), 10504–10512.
6. Reddy, P. V. L., Kavitha, B., Reddy, P. A. K., & Kim, K. H. (2017). TiO2-based photocata-
lytic disinfection of microbes in aqueous media: A review. Environmental Research, 154,
296–303.
7. Reddy, P. V. L., Kim, K. H., & Kim, Y. H. (2011). A review of photocatalytic treatment for
various air pollutants. Asian Journal of Atmospheric Environment, 5(3), 181–188.
8. Reddy, P. A. K., Reddy, P. V. L., Kwon, E., Kim, K. H., Akter, T., & Kalagara, S. (2016).
Recent advances in photocatalytic treatment of pollutants in aqueous media. Environment
International, 91, 94–103.
9. Ozawa, K., Emori, M., Yamamoto, S., Yukawa, R., Yamamoto, S., Hobara, R., Fujikawa, K.,
Sakama, H., & Matsuda, I. (2014). Electron–hole recombination time at TiO2 single-crystal
surfaces: Influence of surface band bending. Journal of Physical Chemistry Letters, 5(11),
1953–1957.
10. Nolan, M., Iwaszuk, A., Lucid, A. K., Carey, J. J., & Fronzi, M. (2016). Design of novel vis-
ible light active photocatalyst materials: Surface modified TiO2. Advanced Materials, 28(27),
5425–5446.
11. Chand, R., Obuchi, E., Katoh, K., Luitel, H. N., & Nakano, K. (2013). Effect of transition
metal doping under reducing calcination atmosphere on photocatalytic property of TiO2
immobilized on SiO2 beads. Journal of Environmental Sciences, 25(7), 1419–1423.
12. Roy, N., Sohn, Y., Leung, K. T., & Pradhan, D. (2014). Engineered electronic states of transi-
tion metal doped TiO2 nanocrystals for low overpotential oxygen evolution reaction. Journal
of Physical Chemistry C, 118(51), 29499–29506.
13. Reddy, P. A. K., Reddy, P. V. L., Sharma, V. M., Srinivas, B., Kumari, V. D., & Subrahmanyam,
M. (2010). Photocatalytic degradation of isoproturon pesticide on C, N and S doped TiO2.
Journal of Water Resource and Protection, 2, 235–244.
14. Vaiano, V., Sacco, O., Sannino, D., Ciambelli, P., Longo, S., Venditto, V., & Guerra, G. (2014).
N-doped TiO2/s-PS aerogels for photocatalytic degradation of organic dyes in wastewater
under visible light irradiation. Journal of Chemical Technology and Biotechnology, 89(8),
1175–1181.
15. D’Souza, L. P., Shwetharani, R., Amoli, V., Fernando, C. A. N., Sinha, A. K., & Balakrishna,
R. G. (2016). Photoexcitation of neodymium doped TiO2 for improved performance in dye-­
sensitized solar cells. Materials and Design, 104, 346–354.
16. Lu, N., Yeh, Y. P., Wang, G. B., Feng, T. Y., Shih, Y. H., & Chen, D. (2017). Dye-sensitized
TiO2-catalyzed photodegradation of sulfamethoxazole under blue or yellow light.
Environmental Science and Pollution Research International, 24(1), 489–499.
17. Lepcha, A., Maccato, C., Mettenbörger, A., Andreu, T., Mayrhofer, L., Walter, M., Olthof, S.,
Ruoko, T.-P., Klein, A., Moseler, M., Meerholz, K., Morante, J. R., Barreca, D., & Mathur,
S. (2015). Electrospun black titania nanofibers: Influence of hydrogen plasma-induced dis-
order on the electronic structure and photoelectrochemical performance. Journal of Physical
Chemistry C, 119(33), 18835–18842.
18. Ullattil, S. G., Narendranath, S. B., Pillai, S. C., & Periyat, P. (2018). Black TiO2 nanomateri-
als: A review of recent advances. Chemical Engineering Journal, 343, 708–736.
19. Chen, X., Liu, L., & Huang, F. (2015). Black titanium dioxide (TiO2) nanomaterials. Chemical
Society Reviews, 44, 1861–1885.
20. Yan, X., Li, Y., & Xia, T. (2017). Black titanium dioxide nanomaterials in photocatalysis.
International Journal of Photoenergy, 8529851, 1–16.
21. Qiu, J., Li, S., Gray, E., Liu, H., Gu, Q.-F., Sun, C., Lai, C., Zhao, H., & Zhang, S. (2014).
Hydrogenation synthesis of blue TiO2 for high-performance lithium-ion batteries. Journal of
Physical Chemistry C, 118(17), 8824–8830.
292 P. A. K. Reddy et al.

22. Liu, N., Schneider, C., Freitag, D., Hartmann, M., Venkatesan, U., Muller, J., Spiecker, E.,
& Schmuki, P. (2014). Black TiO2 nanotubes: Cocatalyst-free open-circuit hydrogen genera-
tion. Nano Letters, 14(6), 3309–3313.
23. Chen, X., Liu, L., Yu, P. Y., & Mao, S. S. (2011). Increasing solar absorption for photocataly-
sis with black hydrogenated titanium dioxide nanocrystals. Science, 331(6018), 746–750.
24. Lu, H., Zhao, B., Pan, R., Yao, J., Qiu, J., Luo, L., & Liu, Y. (2014). Safe and facile hydroge-
nation of commercial Degussa P25 at room temperature with enhanced photocatalytic activ-
ity. RSC Advances, 4, 1128–1132.
25. Liu, Y., Feng, H., Yan, X., Wang, J., Yang, H., Du, Y., & Hao, W. (2017). The origin of
enhanced photocatalytic activities of hydrogenated TiO2 nanoparticles. Dalton Transactions,
46, 10694–10699.
26. Liu, H., Ma, H. T., Li, X. Z., Li, W. Z., Wu, M., & Bao, X. H. (2003). The enhancement of
TiO2 photocatalytic activity by hydrogen thermal treatment. Chemosphere, 50(1), 39–46.
27. Yu, X., Kim, B., & Kim, Y.  K. (2013). Highly enhanced photoactivity of anatase TiO2
nanocrystals by controlled hydrogenation-induced surface defects. ACS Catalysis, 3(11),
2479–2486.
28. Wang, G., Wang, H., Ling, Y., Tang, Y., Yang, X., Fitzmorris, R. C., Wang, C., Zhang, J. Z., &
Li, Y. (2011). Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water split-
ting. Nano Letters, 11(7), 3026–3033.
29. Naldoni, A., Allieta, M., Santangelo, S., Marelli, M., Fabbri, F., Cappelli, S., Bianchi, C. L.,
Psaro, R., & Santo, V. D. (2012). Effect of nature and location of defects on bandgap nar-
rowing in black TiO2 nanoparticles. Journal of the American Chemical Society, 134(18),
7600–7603.
30. Shin, J.-Y., Joo, J. H., Samuelis, D., & Maier, J. (2012). Oxygen-deficient TiO2−δ nanoparti-
cles via hydrogen reduction for high rate capability lithium batteries. Chemistry of Materials,
24(3), 543–551.
31. Danon, A., Bhattacharyya, K., Vijayan, B. K., Lu, J., Sauter, D. J., Gray, K. A., Stair, P. C.,
& Weitz, E. (2012). Effect of reactor materials on the properties of titanium oxide nanotubes.
ACS Catalysis, 2(1), 45–49.
32. Zhang, S., Zhang, S., Peng, B., Wang, H., Yu, H., Wang, H., & Peng, F. (2014). High per-
formance hydrogenated TiO2 nanorod arrays as a photoelectrochemical sensor for organic
compounds under visible light. Electrochemistry Communications, 40, 24–27.
33. Zhu, W.-D., Wang, C.-W., Chen, J.-B., Li, D.-S., Zhou, F., & Zhang, H.-L. (2012). Enhanced
field emission from hydrogenated TiO2 nanotube arrays. Nanotechnology, 23, 455204.
34. He, H., Yang, K., Wang, N., Luo, F., & Chen, H. (2013). Hydrogenated TiO2 film for enhanc-
ing photovoltaic properties of solar cells and self-sensitized effect. Journal of Applied
Physics, 114, 213505.
35. Wang, D., Zhang, X., Sun, P., Lu, S., Wang, L., Wang, C., & Liu, Y. (2014).
Photoelectrochemical water splitting with rutile TiO2 nanowires array: Synergistic effect
of hydrogen treatment and surface modification with anatase nanoparticles. Electrochimica
Acta, 130, 290–295.
36. Zhu, Y., Liu, D., & Meng, M. (2014). H2 spillover enhanced hydrogenation capability of TiO2
used for photocatalytic splitting of water: A traditional phenomenon for new applications.
Chemical Communications, 50, 6049–6051.
37. Zhang, H., Xing, Z., Zhang, Y., Li, Z., Wu, X., Liu, C., Zhu, Q., & Zhou, W. (2015). Ni2+
and Ti3+ co-doped porous black anatase TiO2 with unprecedented-high visible-light-driven
photocatalytic degradation performance. RSC Advances, 5, 107150–107157.
38. Grabstanowicz, L. R., Gao, S., Li, T., Rickard, R. M., Rajh, T., Liu, D. J., & Xu, T. (2013).
Facile oxidative conversion of TiH2 to high-concentration Ti3+-self-doped rutile TiO2 with
visible-light photoactivity. Inorganic Chemistry, 52(7), 3884–3890.
39. Myung, S.  T., Kikuchi, M., Yoon, C.  S., Yashiro, H., Kim, S.  J., Sun, Y.  K., & Scrosati,
B. (2013). Black anatase titania enabling ultra high cycling rates for rechargeable lithium
batteries. Energy & Environmental Science, 6, 2609–2614.
11  Black TiO2: An Emerging Photocatalyst and Its Applications 293

40. Li, L., Chen, Y., Jiao, S., Fang, Z., Liu, X., Xu, Y., Pang, G., & Feng, S. (2016). Synthesis,
microstructure, and properties of black anatase and B phase TiO2 nanoparticles. Materials
and Design, 100, 235–240.
41. Wei, S., Wu, R., Xu, X., Jian, J., Wang, H., & Sun, Y. (2016). One-step synthetic approach
for core-shelled black anatase titania with high visible light photocatalytic performance.
Chemical Engineering Journal, 299, 120–125.
42. Leshuk, T., Parviz, R., Everett, P., Krishnakumar, H., Varin, R.  A., & Gu, F. (2013).
Photocatalytic activity of hydrogenated TiO2. ACS Applied Materials & Interfaces, 5(6),
1892–1895.
43. Wang, Z., Yang, C., Lin, T., Yin, H., Chen, P., Wan, D., Xu, F., Huang, F., Lin, J., Xie, X., &
Jiang, M. (2013). H-doped black titania with very high solar absorption and excellent photo-
catalysis enhanced by localized surface plasmon resonance. Advanced Functional Materials,
23, 5444–5450.
44. Teng, F., Li, M., Gao, C., Zhang, G., Zhang, P., Wang, Y., Chen, L., & Xie, E. (2014).
Preparation of black TiO2 by hydrogen plasma assisted chemical vapor deposition and its
photocatalytic activity. Applied Catalysis B: Environmental, 148–149, 339–343.
45. Yan, Y., Hao, B., Wang, D., Chen, G., Markweg, E., Albrecht, A., & Schaaf, P. (2013).
Understanding the fast lithium storage performance of hydrogenated TiO2 nanoparticles.
Journal of Materials Chemistry A, 1, 14507–14513.
46. Panomsuwan, G., Watthanaphanit, A., Ishizaki, T., & Saito, N. (2015). Water-plasma-assisted
synthesis of black titania spheres with efficient visible-light photocatalytic activity. Physical
Chemistry Chemical Physics, 17, 13794–13799.
47. Wang, Z., Yang, C., Lin, T., Yin, H., Chen, P., Wan, D., Xu, F., Huang, F., Lin, J., Xie, X., &
Jiang, M. (2013). Visible-light photocatalytic, solar thermal and photoelectrochemical prop-
erties of aluminium-reduced black titania. Energy & Environmental Science, 6, 3007–3014.
48. Cui, H., Zhao, W., Yang, C., Yin, H., Lin, T., Shan, Y., & Huang, F. (2014). Black TiO2 nano-
tube arrays for high-efficiency photoelectrochemical water-splitting. Journal of Materials
Chemistry A, 2, 8612–8616.
49. Lin, T., Yang, C., Wang, Z., Yin, H., Lü, X., Huang, F., & Jiang, M. (2014). Effective
nonmetal incorporation in black titania with enhanced solar energy utilization. Energy &
Environmental Science, 7, 967–972.
50. Zhao, Z., Tan, H., Zhao, H., Lv, Y., Zhou, L.-J., Song, Y., & Sun, Z. (2014). Reduced TiO2
rutile nanorods with well-defined facets and their visible-light photocatalytic activity.
Chemical Communications, 50, 2755–2757.
51. Sinhamahapatra, A., Jeon, J.  P., & Yu, J.  S. (2015). A new approach to prepare highly
active and stable black titania for visible light-assisted hydrogen production. Energy &
Environmental Science, 8, 3539–3544.
52. Zhang, K., Wang, L., Kim, J.  K., Ma, M., Veerappan, G., Lee, C.-L., Kong, K.-J., Lee,
H., & Park, J.  H. (2016). An order/disorder/water junction system for highly efficient
co-catalyst-­free photocatalytic hydrogen generation. Energy & Environmental Science, 9,
499–503.
53. Kang, Q., Cao, J., Zhang, Y., Liu, L., Xu, H., & Ye, J. (2013). Reduced TiO2 nanotube arrays
for photoelectrochemical water splitting. Journal of Materials Chemistry A, 1, 5766–5774.
54. Tan, H., Zhao, Z., Niu, M., Mao, C., Cao, D., Cheng, D., Feng, P., & Sun, Z. (2014). A facile
and versatile method for preparation of colored TiO2 with enhanced solar-driven photocata-
lytic activity. Nanoscale, 6, 10216–10223.
55. Tominaka, S., Tsujimoto, Y., Matsushita, Y., & Yamaura, K. (2011). Synthesis of nanostruc-
tured reduced titanium oxide: Crystal structure transformation maintaining nanomorphology.
Angewandte Chemie International Edition, 50, 7418–7421.
56. Zhu, G., Yin, H., Yang, C., Cui, H., Wang, Z., Xu, J., Lin, T., & Huang, F. (2015). Black titania
for superior photocatalytic hydrogen production and photoelectrochemical water splitting.
ChemCatChem, 7, 2614–2619.
294 P. A. K. Reddy et al.

57. Zhang, Z., Hedhili, M. N., Zhu, H., & Wang, P. (2013). Electrochemical reduction induced
self-doping of Ti3+ for efficient water splitting performance on TiO2 based photoelectrodes.
Physical Chemistry Chemical Physics, 15, 15637–15644.
58. Zheng, L., Cheng, H., Liang, F., Shu, S., Tsang, C. K., Li, H., & Li, Y. Y. (2012). Porous
TiO2 photonic band gap materials by anodization. Journal of Physical Chemistry C, 116(9),
5509–5515.
59. Dong, J., Han, J., Liu, Y., Nakajima, A., Matsushita, S., Wei, S., & Gao, W. (2014). Defective
black TiO2 synthesized via anodization for visible-light photocatalysis. ACS Applied
Materials & Interfaces, 6(3), 1385–1388.
60. Chen, X., Zhao, D., Liu, K., Wang, C., Liu, L., Li, B., & Shen, D. (2015). Laser-modified
black titanium oxide nanospheres and their photocatalytic activities under visible light. ACS
Applied Materials & Interfaces, 7, 16070–16077.
61. Nakajima, T., Nakamura, T., Shinoda, K., & Tsuchiya, T. (2014). Rapid formation of black
titania photoanodes: Pulsed laser-induced oxygen release and enhanced solar water splitting
efficiency. Journal of Materials Chemistry A, 2, 6762–6771.
62. Lü, X., Chen, A., Luo, Y., Lu, P., Dai, Y., Enriquez, E., Dowden, P., Xu, H., Kotula, P. G.,
Azad, A.  K., Yarotski, D.  A., Prasankumar, R.  P., Taylor, A.  J., Thompson, J.  D., & Jia,
Q. (2016). Conducting interface in oxide homojunction: Understanding of superior proper-
ties in black TiO2. Nano Letters, 16(9), 5751–5755.
63. Li, G., Lian, Z., Li, X., Xu, Y., Wang, W., Zhang, D., Tian, F., & Li, H. (2015). Ionothermal
synthesis of black Ti3+-doped single-crystal TiO2 as an active photocatalyst for pollutant deg-
radation and H2 generation. Journal of Materials Chemistry A, 3, 3748–3756.
64. Zou, X., Liu, J., Su, J., Zuo, F., Chen, J., & Feng, P. (2013). Facile synthesis of thermal-
and photostable titania with paramagnetic oxygen vacancies for visible-light photocatalysis.
Chemistry—A European Journal, 19, 2866–2873.
65. Liu, N., Häublein, V., Zhou, X., Venkatesan, U., Hartmann, M., Mačković, M., Nakajima, T.,
Spiecker, E., Osvet, A., Frey, L., & Schmuki, P. (2015). “Black” TiO2 nanotubes formed by
high-energy proton implantation show noblemetal-co-catalyst free photocatalytic H2−evolu-
tion. Nano Letters, 15(10), 6815–6820.
66. Huang, H., Zhang, H., Ma, Z., Liu, Y., Zhang, X., Han, Y., & Kang, Z. (2013). Si quantum
dot-assisted synthesis of mesoporous black TiO2 nanocrystals with high photocatalytic activ-
ity. Journal of Materials Chemistry A, 1, 4162–4166.
67. Xia, T., & Chen, X. (2013). Revealing the structural properties of hydrogenated black TiO2
nanocrystals. Journal of Materials Chemistry A, 1, 2983–2989.
68. Wang, W., Ni, Y., Lu, C., & Xu, Z. (2012). Hydrogenation of TiO2 nanosheets with exposed
{001} facets for enhanced photocatalytic activity. RSC Advances, 2, 8286–8288.
69. Jiang, X., Zhang, Y., Jiang, J., Rong, Y., Wang, Y., Wu, Y., & Pan, C. (2012). Characterization
of oxygen vacancy associates within hydrogenated TiO2: A positron annihilation study.
Journal of Physical Chemistry C, 116(42), 22619–22624.
70. Lu, Z., Yip, C.-T., Wang, L., Huang, H., & Zhou, L. (2012). Hydrogenated TiO2 nanotube
arrays as high-rate anodes for lithium-ion microbatteries. ChemPlusChem, 77, 991–1000.
71. Zeng, L., Song, W., Li, M., Zeng, D., & Xie, C. (2014). Catalytic oxidation of formalde-
hyde on surface of H-TiO2/H-C-TiO2 without light illumination at room temperature. Applied
Catalysis. B, Environmental, 147, 490–498.
72. Chen, X., Liu, L., Liu, Z., Marcus, M. A., Wang, W.-C., Oyler, N. A., Grass, M. E., Mao, B.,
Glans, P.-A., Yu, P. Y., Guo, J., & Mao, S. S. (2013). Properties of disorder engineered black
titanium dioxide nanoparticles through hydrogenation. Scientific Reports, 3, 1510.
73. Xia, T., Zhang, C., Oyler, N.  A., & Chen, X. (2013). Hydrogenated TiO2 nanocrystals: A
novel microwave absorbing material. Advanced Materials, 25, 6905–6910.
74. Zhang, C., Yu, H., Li, Y., Gao, Y., Zhao, Y., Song, W., Shao, Z., & Yi, B. (2013). Supported
Noble metals on hydrogen-treated TiO2 nanotube arrays as highly ordered electrodes for fuel
cells. ChemSusChem, 6, 659–666.
11  Black TiO2: An Emerging Photocatalyst and Its Applications 295

75. Xia, T., Zhang, C., Oyler, N. A., & Chen, X. (2014). Enhancing microwave absorption of
TiO2 nanocrystals via hydrogenation. Journal of Materials Research, 29, 2198–2210.
76. Lu, X., Wang, G., Zhai, T., Yu, M., Gan, J., Tong, Y., & Li, Y. (2012). Hydrogenated TiO2
nanotube arrays for supercapacitors. Nano Letters, 12(3), 1690–1696.
77. Yin, H., Lin, T., Yang, C., Wang, Z., Zhu, G., Xu, T., Xie, X., Huang, F., & Jiang, M. (2013).
Gray TiO2 nanowires synthesized by aluminum-mediated reduction and their excellent pho-
tocatalytic activity for water cleaning. Chemistry—A European Journal, 19, 13313–13316.
78. Rekoske, J. E., & Barteau, M. A. (1997). Isothermal reduction kinetics of titanium dioxide-­
based materials. The Journal of Physical Chemistry. B, 101(7), 1113–1124.
79. Li, S., Qiu, J., Ling, M., Peng, F., Wood, B., & Zhang, S. (2013). Photoelectrochemical char-
acterization of hydrogenated TiO2 nanotubes as photoanodes for sensing applications. ACS
Applied Materials & Interfaces, 5(21), 11129–11135.
80. Pesci, F. M., Wang, G. M., Klug, D. R., Li, Y., & Cowan, A. J. (2013). Efficient suppression of
electron–hole recombination in oxygen-deficient hydrogen-treated TiO2 nanowires for photo-
electrochemical water splitting. Journal of Physical Chemistry C, 117(48), 25837–25844.
81. Zhu, G., Lin, T., Lu, X., Zhao, W., Yang, C., Wang, Z., Yin, H., Liu, Z., Huang, F., & Lin,
J. (2013). Black brookite titania with high solar absorption and excellent photocatalytic per-
formance. Journal of Materials Chemistry A, 1, 9650–9653.
82. Pei, Z., Ding, L., Lin, H., Weng, S., Zheng, Z., Hou, Y., & Liu, P. (2013). Facile synthe-
sis of defect-mediated TiO2−x with enhanced visible light photocatalytic activity. Journal of
Materials Chemistry A, 1, 10099–10102.
83. Qiu, J., Lai, C., Gray, E., Li, S., Qiu, S., Strounina, E., Sun, C., Zhao, H., & Zhang, S. (2014).
Blue hydrogenated lithium titanate as a high-rate anode material for lithium-ion batteries.
Journal of Materials Chemistry A, 2, 6353–6358.
84. Zheng, Z., Huang, B., Lu, J., Wang, Z., Qin, X., Zhang, X., Dai, Y., & Whangbo, M.-H. (2012).
Hydrogenated titania: Synergy of surface modification and morphology improvement for
enhanced photocatalytic activity. Chemical Communications, 48, 5733–5735.
85. Tian, M., Samani, M.  M., Eres, G., Sachan, R., Yoon, M., Chisholm, M.  F., Wang, K.,
Puretzky, A. A., Rouleau, C. M., Geohegan, D. B., & Duscher, G. (2015). Structure and for-
mation mechanism of black TiO2 nanoparticles. ACS Nano, 9(10), 10482–10488.
86. Zuo, F., Wang, L., Wu, T., Zhang, Z., Borchardt, D., & Feng, P. (2010). Self-doped Ti3+
enhanced photocatalyst for hydrogen production under visible light. Journal of the American
Chemical Society, 132(34), 11856–11857.
87. Wang, J., Zhang, P., Li, X., Zhu, J., & Li, H. (2013). Synchronical pollutant degradation and
H2 production on a Ti3+-doped TiO2 visible photocatalyst with dominant (001) facets. Applied
Catalysis B: Environmental, 134–135, 198–204.
88. Zhou, Y., Chen, C., Wang, N., Li, Y., & Ding, H. (2016). Stable Ti3+ self-doped anatase-rutile
mixed TiO2 with enhanced visible light utilization and durability. The Journal of Physical
Chemistry C, 120, 6116–6124.
89. Zhang, X., Wang, J., Hu, W., Zhang, K., Sun, B., Tian, G., Jiang, B., Pan, K., & Zhou,
W. (2016). Facile strategy to fabricate uniform black TiO2 nanothorns/graphene/black TiO2
nanothorns sandwich like nanosheets for excellent solar-driven photocatalytic performance.
ChemCatChem, 8(20), 3240–3246.
90. Liu, X., Xing, Z., Zhang, H., Wang, W., Yan, Z., Li, Z., Wu, X., Yu, X., & Zhou, W. (2016).
Fabrication of 3D mesoporous black TiO2/MoS2/TiO2 nanosheets for visible-light-driven
photocatalysis. ChemSusChem, 9(10), 1118–1124.
91. Serra, M., Khan, A., Asihi, A. M., Kosa, S. A., & Garcia, H. (2015). Photocatalytic hydrogen
generation from water–methanol mixtures using “black” anatase obtained by annealing of
titanate nanotubes. Materials Today Communications, 4, 63–68.
92. Hu, W., Zhou, W., Zhang, K., Zhang, X., Wang, L., Jiang, B., Tian, G., Zhao, D., & Fu,
H. (2016). Facile strategy for controllable synthesis of stable mesoporous black TiO2 hollow
spheres with efficient solar-driven photocatalytic hydrogen evolution. Journal of Materials
Chemistry A, 4, 7495–7502.
296 P. A. K. Reddy et al.

93. Zhou, W., Li, W., Wang, J.-Q., Qu, Y., Yang, Y., Xie, Y., Zhang, K., Wang, L., Fu, H., & Zhao,
D. (2014). Ordered mesoporous black TiO2 as highly efficient hydrogen evolution photocata-
lyst. Journal of the American Chemical Society, 136(26), 9280–9283.
94. Xin, X., Xu, T., Wang, L., & Wang, C. (2016). Ti3+- self doped brookite TiO2 single-­crystalline
nanosheets with high solar absorption and excellent photocatalytic CO2 reduction. Scientific
Reports, 6, 23684.
95. Qingli, W., Zhaoguo, Z., Xudong, C., Zhengfeng, H., Peimei, D., Yi, C., & Xiwen, Z. (2015).
Photoreduction of CO2 using black TiO2 films under solar light. Journal of CO2 Utilization,
12, 7–11.
96. Zhao, J., Li, Y., Zhu, Y., Wang, Y., & Wang, C. (2016). Enhanced CO2 photoreduction activity
of black TiO2−coated Cu nanoparticles under visible light irradiation: Role of metallic Cu.
Applied Catalysis A: General, 510, 34–41.
97. Han, B., Wei, W., Chang, L., Cheng, P., & Hu, Y. H. (2016). Efficient visible light photocata-
lytic CO2 reforming of CH4. ACS Catalysis, 6(2), 494–497.
98. Tian, L., Xu, J., Just, M., Green, M., Liu, L., & Chen, X. (2017). Broad range energy absorp-
tion enabled by hydrogenated TiO2 nanosheets: From optical to infrared and microwave.
Journal of Materials Chemistry C, 5, 4645–4653.
99. Zhang, K., Zhou, W., Zhang, X., Qu, Y., Wang, L., Hu, W., Pan, K., Li, M., Xie, Y., Jiang,
B., & Tian, G. (2016). Large-scale synthesis of stable mesoporous black TiO2 nanosheets
for efficient solar-driven photocatalytic hydrogen evolution via an earth-abundant low-cost
biotemplate. RSC Advances, 6, 50506–50512.
100. Leshuk, T., Linley, S., & Gu, F. (2013). Hydrogenation processing of TiO2 nanoparticles.
Canadian Journal of Chemical Engineering, 91, 799–807.
101. Wu, M. C., Chang, I. C., Hsiao, K. C., & Huang, W. K. (2016). Highly visible-light absorbing
black TiO2 nanocrystals synthesized by sol–gel method and subsequent heat treatment in low
partial pressure H2. Journal of the Taiwan Institute of Chemical Engineers, 63, 430–435.
102. Han, L., Ma, Z., Luo, Z., Liu, G., Ma, J., & An, X. (2016). Enhanced visible light and pho-
tocatalytic performance of TiO2 nanotubes by hydrogenation at lower temperature. RSC
Advances, 6, 6643–6650.
103. Yan, Y., Han, M., Konkin, A., Koppe, T., Wang, D., Andreu, T., Chen, G., Vetter, U., Morante,
J. R., & Schaaf, P. (2014). Slightly hydrogenated TiO2 with enhanced photocatalytic perfor-
mance. Journal of Materials Chemistry A, 2, 12708–12716.
104. An, H.-R., Park, S. Y., Kim, H., Lee, C. Y., Choi, S., Lee, S. C., Seo, S., Park, E. C., Oh,
Y.-K., Song, C.-G., Won, J., Kim, Y. J., Lee, J., Lee, H. U., & Lee, Y.-C. (2016). Advanced
nanoporous TiO2 photocatalysts by hydrogen plasma for efficient solar-light photocatalytic
application. Scientific Reports, 6, 29683.
105. Tian, J., Hu, X., Yang, H., Zhou, Y., Cui, H., & Liu, H. (2016). High yield production of
reduced TiO2 with enhanced photocatalytic activity. Applied Surface Science, 360, 738–743.
106. Yan, B., Zhou, P., Xu, Q., Zhou, X., Xu, D., & Zhu, J. (2016). Engineering disorder into exotic
electronic 2D TiO2 nanosheets for enhanced photocatalytic performance. RSC Advances, 6,
6133–6137.
107. Ye, M., Jia, J., Wu, Z., Qian, C., Chen, R., O’Brien, P.  G., Sun, W., Dong, Y., & Ozin,
G. A. (2017). Synthesis of black TiOx nanoparticles by mg reduction of TiO2 nanocrystals
and their application for solar water evaporation. Advanced Energy Materials, 7, 1601811.
108. Xu, C., Song, Y., Lu, L., Cheng, C., Liu, D., Fang, X., & Li, D. (2013). Electrochemically
hydrogenated TiO2 nanotubes with improved photoelectrochemical water splitting perfor-
mance. Nanoscale Research Letters, 8, 391.
109. Li, H., Chen, Z., Tsang, C. K., Li, Z., Ran, X., Lee, C., & Pan, B. (2014). Electrochemical
doping of anatase TiO2 in organic electrolytes for high-performance supercapacitors and pho-
tocatalysts. Journal of Materials Chemistry A, 2, 229–236.
110. Han, K., Zhang, X., Wang, H., Liu, Y., & Cao, A. (2016). A facile microwaving method to
turn titanium oxide into highly active Ti3+ self-doped structure. Journal of Nanoscience and
Nanotechnology, 16, 9826–9831.
11  Black TiO2: An Emerging Photocatalyst and Its Applications 297

111. Liu, X., Xing, Z., Zhang, Y., Li, Z., Wu, X., Tan, S., Yu, X., Zhu, Q., & Zhou, W. (2017).
Fabrication of 3D flower-like black N-TiO2-x@MoS2 for unprecedented-high visible-light-­
driven photocatalytic performance. Applied Catalysis. B, Environmental, 201, 119–127.
112. Fan, C., Fua, X., Shia, L., Yua, S., Qiana, G., & Wanga, Z. (2016). Ultrasonic-induced nano-
composites with anatase@amorphous TiO2 core–shell structure and their photocatalytic
activity. RSC Advances, 6, 67444–67448.
113. Tian, Z., Cui, H., Zhu, G., Zhao, W., Xu, J., Shao, F., He, J., & Huang, F. (2016). Hydrogen
plasma reduced black TiO2-B nanowires for enhanced photoelectrochemical water-splitting.
Journal of Power Sources, 325, 697–705.
114. Wang, C.-C., & Chou, P.-H. (2016). Effects of various hydrogenated treatments on formation
and photocatalytic activity of black TiO2 nanowire arrays. Nanotechnology, 27, 325401.
115. Xu, J., Tian, Z., Yin, G., Lin, T., & Huang, F. (2017). Controllable reduced black titania
with enhanced photoelectrochemical water splitting performance. Dalton Transactions, 46,
1047–1051.
116. Wang, H., Lin, T., Zhu, G., Yin, H., Lü, X., Li, Y., & Huang, F. (2015). Colored titania
nanocrystals and excellent photocatalysis for water cleaning. Catalysis Communications,
60, 55–59.
117. Xin, X., Xu, T., Yin, J., Wang, L., & Wang, C. (2015). Management on the location and
concentration of Ti3+ in anatase TiO2 for defects-induced visible-light photocatalysis. Applied
Catalysis B: Environmental, 176, 354–362.
118. Xin, L., & Liu, X. (2015). Black TiO2 inverse opals for visible-light photocatalysis. RSC
Advances, 5, 71547–71550.
119. Samsudin, E.  M., Hamid, S.  B. A., Juan, J.  C., Basirun, W.  J., & Kandjani, A.  E. (2015).
Surface modification of mixed-phase hydrogenated TiO2 and corresponding photocatalytic
response. Applied Surface Science, 359, 883–896.
120. Yang, C., Wang, Z., Lin, T., Yin, H., Lü, X., Wan, D., & Xie, X. (2013). Core-shell nanostruc-
tured “black” rutile titania as excellent catalyst for hydrogen production enhanced by sulfur
doping. Journal of the American Chemical Society, 135(47), 17831–17838.
121. Kim, Y., Hwang, H. M., Wang, L., Kim, I., Yoon, Y., & Lee, H. (2016). Solar-light photocata-
lytic disinfection using crystalline/amorphous low energy bandgap reduced TiO2. Scientific
Reports, 6, 25212.
Chapter 12
Nanomaterials for Photocatalytic
Decomposition of Endocrine Disruptors
in Water

Ajay Kumar, Vishal Sharma, Ashish Kumar, and Venkata Krishnan

12.1  Introduction

Pure water is essential for sustaining life at the earth as the consumption of contami-
nated water causes serious ill-health issues worldwide [1]. According to the World
Health Organization (WHO), approximately 2.1 billion people do not have access
to pure drinking water and almost double as many lack basic sanitation [2]. Water is
considered to be pure when it is colorless, non-turbid, and without abnormal smell
and taste. On the other hand, it is polluted when it contains poisonous chemical
substances, domestic and industrial sewage, and organic and inorganic substances
[3]. Several chemicals are used in our daily life, but their release and long-term
accumulation in water bodies can cause serious harm to the environment [4]. Many
uncontrolled chemicals released into the environment cause adverse biological
effects for animals and humans. Common examples are chlorinated organic com-
pounds arising from sewage water or chemical plants, certain compounds used in
agriculture, textile dyes, pollutants emanating from mining activity, and other exter-
nalities of various manufacturing activities [1, 5–7]. Most of them are endocrine
disrupting chemicals (EDC), whose concentration has become uncontrolled in the
water bodies and their regulation is a global challenge [8–11]. Inside the human
body, there are network of organs and glands that produce, store, and secrete hor-
mones known as endocrine system. Normally, these hormones travel from one gland
or organ to another and fit into specific receptors there. When the EDC enter inside
the body, they can block the hormones from fitting into the receptor or fit into the
receptor itself. This results in abnormal processes in the body or sometimes severe

Ajay Kumar and Vishal Sharma contributed equally to this work.

A. Kumar · V. Sharma · A. Kumar · V. Krishnan (*)


School of Basic Sciences and Advanced Materials Research Center, Indian Institute of
Technology Mandi, Kamand, Mandi, Himachal Pradesh, India
e-mail: vkn@iitmandi.ac.in

© Springer Nature Switzerland AG 2021 299


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_12
300 A. Kumar et al.

health consequences [12]. The concept of endocrine disruptor was first introduced
in World Wildlife Fund Wingspread Conference in 1991 [13]. Further work in this
direction led American Endocrine Society to propose the mechanism of endocrine
disruption [14]. After that, extensive research work has been started to understand
the health risk caused by EDC.
EDC are the harmonically active environmental pollutants that can interfere with
endocrine systems in human or animal bodies. The presence of an excessive number
of EDC is responsible for various health-related issues such as heart problems,
growth and obesity, abnormalities in sex organs, immune system, cancerous tumor,
nervous system function, neurological and learning disabilities, and many more
because almost every process in our body are operated by hormones [15]. EDC are
ubiquitous as they are in industrial chemicals, pesticides, fungicides, etc. from
where they leach into soil and ground water and enter into the food chain. Also, they
are present in consumer products such as plastics, fabrics, cosmetics, lotions, per-
fumes and antibacterial soaps, etc. EDC as a target substrates for decomposition can
be categorized into five types: (1) pesticides and herbicides, (2) hormones and ste-
roids, (3) plasticizers (bisphenols and phthalates), (4) additives of personal care
products (phenones, parabens etc.), and (5) organic compounds such as polybromi-
nated diphenyl ethers (PBDE), perfluorinated chemicals (PFC), carbamazepine,
metformin, phenols, dioxins, polycyclic aromatic hydrocarbons (PAH), etc. [9, 16].
For the decomposition of EDC in waste water bodies, various processes have been
employed such as ultrasonic [17], biological [18], adsorption [19, 20], and homoge-
neous photochemical reactions [21]. Even though these methods are helpful for the
removal of EDC from waste water, still there are many shortcomings in various
aspects. The physical treatment of EDC only includes their transfer from one
medium to another without their subsequent decomposition, whereas the biological
treatment requires crucial efforts to cultivate the bacteria and EDC decomposition
rate is also low as compare to other methods. Of course, homogeneous photocataly-
sis offers poor recyclability of the catalyst used in the treatment. However, the use
of heterogeneous photocatalysis is the most effective and convenient approach for
decomposition of EDC due to several reasons such as complete mineralization and
removal of most of pollutants, low-energy consumption, and mild reactions condi-
tions [22, 23].
When semiconducting nanomaterials are illuminated with light energy more
than the bandgap, the electron jumps from the valence band (VB) to the conduction
band (CB) leaving the hole behind in the VB thereby separating charge carriers [24,
25]. These separated charge carriers than start migrating towards the surface of
nanomaterial where they under oxidation and reduction reactions with adsorbed
water and oxygen molecules results in the formation of O2•− and OH• radicals,
respectively [26, 27]. These generated active species further initiate the photocata-
lytic reaction by attacking the pollutant molecules. Usually, the charge carriers limit
the photocatalytic activity because of their fast recombination rate. The rate of
transfer of charge carriers depends upon the band edge positions of nanomaterial
composites and on redox potential of adsorbed species [28]. There are various strat-
egies used to enhance the photocatalytic activity of nanomaterials such as
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 301

combination of two semiconductors [29–34], doping with metals [35], and non-
metals [36] by separating the charge carriers. The determination of band gap values
and the positions of band edges are important to design efficient photocatalyst.
Among the various semiconductor materials, most of them exhibit the photocata-
lytic activity under UV and visible light; however, utilization of only UV/visible
light is not sufficient. Few photocatalysts show maximum wavelength absorption in
near-infrared region directly such as PdO and CuO due to their low band gap and are
used for photocatalytic decompositions [37]. Also, the formation of heterojunction
in nanocomposite semiconductor materials results in enhanced photocatalytic effi-
ciency [38].
In this book chapter, photocatalytic decomposition of various endocrine disrup-
tors using single, binary, ternary, and multi-components nanomaterials has been
discussed in detail. Various photocatalytic nanomaterials, endocrine parts in the
human body, and molecular structures of various endocrine disruptors are shown in
Scheme 12.1. Recent advances in heterogeneous photocatalysis for decompositions

Scheme 12.1  Schematic illustration of the presence of photocatalytic nanomaterials, endocrine


glands in the human body, and molecular structures of various endocrine disruptors
302 A. Kumar et al.

of hazardous EDC are summarized in this book chapter. Furthermore, the degrada-
tion mechanism, different reaction pathways, and the intermediates formed and ana-
lyzed are described in detail.

12.2  Single Component or Unary Nanostructures

The use of single component photocatalysts has been demonstrated by the research-
ers for the decomposition of different EDC. The benchmark TiO2 photocatalyst has
been studied extensively in this regard [39, 40]. TiO2, also known as titania, in
which titanium exist in +4 oxidation state, is a commonly known oxide of titanium.
TiO2 occurs in three different forms in nature: (1) anatase, (2) rutile, and (3) brook-
ite. For example, for the decomposition of endocrine disruptor, resorcinol, Al-Hajji
and coworkers [41] used the mixture of brookite and anatase TiO2 nanowires. The
superior photocatalytic activity of the photocatalyst is due to the interfacial charge
carrier movement between anatase and brookite phases (Fig. 12.1).
Furthermore, Ding et al. [42] have reported fabrication of TiO2 pillared montmo-
rillonites with the help of introducing Ti4+ ions over a layer of montmorillonite
improved with cetyltrimethylammonium bromide (CTAB) to inversigate the photo-
catalytic decomposition of dimethyl phthalate (DMP) under UV light irradiations.
Gas chromatography mass spectrometry was used to determine the intermediates
formed during the photocatalytic decomposition of DMP.  The photoexcitation of
TiO2 and the radicals generated during photocatalytic decomposition are given in
Eqs. 12.1, 12.2, and 12.3, respectively.

TiO2 + hv → e − + h + (12.1)

h + + H 2 O → OH • + H + (12.2)

e − + O2 → O2−• (12.3)

The formation of OH• in the photocatalytic decomposition was verified by the


electron spin resonance spectroscopy in which fourfold peaks with intensity ratio
1:2:2:1 were observed for the DMPO-OH• adduct confirming the formation of OH•
radicals under UV irradiation.
In another report, Xu et al. [43] have utilized the TiO2 as a photocatalyst to inves-
tigate the decomposition of n-butyl benzyl phthalate (BBP) in the presence of UV
light. GC-MS technique was employed to determine the intermediates formed in the
photocatalytic reaction. Based on experimental findings, the photocatalytic decom-
position pathway was given as presented in Fig. 12.2. The results in their findings
revealed that the major intermediates produced in the photocatalytic decomposition
reaction of BBP were monobenzyl phthalate, monobutyl phthalate, and phthalic
acid. Generation of these intermediates results from the donation of the electrons
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 303

Fig. 12.1  Photodegradation mechanism of resorcinol under UV illumination using mixture of


anatase and brookite. Reproduced with permission from [41]. Copyright 2019, Elsevier Publishers

from catalyst and reaction with OH• radical in photocatalytic process. In the first
step, BBP species react with electron and result in the formation of a radical anion,
which can react with OH• and give rise to the anionic product. Loss of benzyl alco-
hol, resultant species were transformed into monobenzyl phthalate and monobutyl
phthalate. Subsequently, reaction with electron and OH• radical, both the species
ended up with the formation of phthalic acid.
Muneer and coworkers [44] have investigated the various intermediates and
degradaed product of 1,2-diethyl phthalate (DEP) by using GC-MS technique and
TiO2 as a photocatalyst. A proposed mechanism for the DEP decomposition has
been presented in Fig. 12.3. The GC-MS measurement of DEP after 195 min results
in the generation of three products such as C8H4O3 (7), C12H16O5 (4), and C10H10O4
(6), which appeared at reaction times of 15, 17, and 27 min, respectively.
In addition to TiO2, ZnO has also been used as a photocatalytic material for the
EDC decomposition [45–49]. Sin and coworkers [50] published the formation of
self-assembly of ZnO hierarchical nano-/microspheres (ZHN) by using surfactant-­
chemical free route. The photocatalytic efficiency of the ZnO hierarchical micro-/
nanospheres was investigated by degrading different EDC such as resorcinol,
bisphenol A, phenol, and methylparaben under UV light. The ZHN showed better
results toward the photocatalytic decomposition of phenol in comparison to bare
ZnO and TiO2. The high surface area and porous hierarchical surface structure pho-
tocatalyst results in the better adsorption of molecules, inhibition of charge carrier
recombination and increased concentration of generated hydroxyl radicals, which
leads to the improved decomposition of EDC. The reaction mechanism for photo-
catalytic degradation of EDC on the surface of ZHN is given by the Eqs. 12.4–12.11
along with the photocatalytic degradation mechanism as shown in Fig. 12.4.



ZnO + hv → ZnO e CB ( +
+ h VB ) (12.4)

h +VB + H 2 O → OH • + H + (12.5)

h +VB + OH − → OH • (12.6)
304 A. Kumar et al.

Fig. 12.2  The proposed


degradation pathways of
BBP in the photocatalytic
reaction over TiO2.
Reproduced with
permission from [43].
Copyright 2008, Elsevier
Publishers
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 305

Fig. 12.3  The proposed degradation pathways of DEP in photocatalytic reaction. Reproduced
with permission from [44]. Copyright 2001, Elsevier Publishers

O2
O2•- • OH
e-

Conduction band degradation


hv e- e- e- e-
products

h+ h+ h+ h+
Valence band h+ H+ + • OH EDC

ZnO
H2O

Fig. 12.4  The mechanism of photocatalytic dedecomposition of EDC over surface of ZHNs.
Reproduced with permission from [50]. Copyright 2013, Elsevier Publishers
306 A. Kumar et al.


e CB + O2 → O2−• (12.7)

O2−• + H + → HOO• (12.8)

HOO• + H + + e CB

→ H 2 O2 (12.9)

H 2 O2 + e CB → OH • + OH − (12.10)

EDC + OH· → Intermediate products → mineralized products (12.11)

Bechambi et al. [51] prepared C-doped ZnO photocatalyst with different amount
of carbon doping using hydrothermal synthesis route. Photocatalytic activity of the
C-doped ZnO was investigated by the decomposition of Bisphenol A (BPA) under
UV light. The observation from their findings revealed that doped ZnO with 4%
carbon content showed best catalytic activity toward photodegradation of BPA
among all the catalysts. The best photocatalytic performance with 4% C-doped ZnO
was ascribed to the effective charge separation and transfer in the optimized cata-
lyst. Furthermore, under optimized conditions, the representative photocatalyst
showed 100% photodecomposition of BPA under UV light irradiation.

12.3  Two Component or Binary Nanostructures

Coupling of two semiconductor materials [8, 52–60] or noble metal deposition [61,
62] can be used to develop more effective catalyst toward the photodegradation of
EDC because the combination of two semiconductors can result in better electron-­
hole pair separation and can result in the improved light absorption. Direct band gap
value of commonly used semiconductor material has been presented in Fig. 12.5,
which can offer a theoretical basis for the development of binary nanostructure pho-
tocatalytic materials [16].
Xu and coworkers [63] published the fabrication of numerious polyoxotungstate/
TiO2 photocatalysts via using combination of sol-gel chemistry and solvothermal
method. The photocatalytic performance of the nanocomposites was determined by
degrading dibutyl phthalate (DBP) under the simulated solar irradiations. In addi-
tion to this, electrospray ionization mass spectrometry (ESI-MS) with ion chroma-
tography analysis was used for the investigation of photocatalytic intermediates.
The decomposed intermediates produced during the photocatalytic process and the
decomposition pathway are presented in Fig. 12.6. In the DBP/TiO2/PW12 system
in the presence of sunlight, there are two pathways, in which OH• radicals can react
with (1) aliphatic chain and (2) aromatic ring which corresponds to DBP molecule.
As shown in Fig. 12.6, reaction of OH• with the aromatic ring results in the genera-
tion of hydroxylated intermediate products (1 and 2), whereas OH• can react with
aliphatic chain of DBP molecule to generate another intermediate product (3). On
further reaction with OH• radicals, intermediate products 1 and 2 loose butoxy
groups and result in the formation of intermediate products 5 and 6. Intermediate
product 3 was again reacted with OH• and led to the generation of intermediate
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 307

Fig. 12.5  Band gap structure of different semiconductor photocatalysts. Reproduced with permis-
sion from [16]. Copyright 2020, Elsevier Publishers

product 7. Intermediate product 4 was generated through alkyl-oxygen bond cleav-


age of DBP molecule. Decarboxylation from intermediate products 4–6 results in
the formation of intermediate products 8 and 9. The alkyl-oxygen bond cleavage in
intermediate product 8 also resulted in intermediate product 9. Further decarboxyl-
ation of compound 9 results in the formation of quinone and dihydroxybenzene.
Finally, quinone was again oxidized, and therefore ring-opening reaction results in
the formation of aliphatic acids such as formic acid, butanedioic acid, and acetic
acid. In addition to this, •OH reacts directly with two C atoms at α-position of DBP
and gives rise to intermediate product 10. All these intermediates ultimately get
oxidized into final products, CO2 and H2O.
Shankar and coworkers have achieved complete mineralization of various ECDs
such as monocrotophos (MCP) and 2,4-dichlorophenoxyacetic acid (DPA) using
TiO2/zeolite binary nanocomposite [64–66]. Khavar et al. [67] have investigated the
photocatalytic efficiency of In,S/TiO2@RGO photocatalyst toward degradation and
detoxification of pesticide atrazine in water under visible irradiation. Their findings
showed that the In, S-TiO2@rGO nanocomposite results in 80 times higher atrazine
degradation rate as compared to TiO2. Moreover, 100% photocatalytic decomposi-
tion and high decomposition of atrazine were achieved using the optimized compos-
ite (with 1 mol% of S, 3 mol% of In and 5 wt.% of rGO) within 20 min. Nawaz and
coworkers [68] reported the formation of TiO2-RGO (RGTO) aerogel photocatalyst
by using the hydrothermal method. The photocatalytic activity was determined by
performing the decomposition of carbamazepine (CBZ) in the presence of UV irra-
diation (Fig. 12.7). The results revealed that the higher photocatalytic performance
was achieved in case of RGTO aerogel than that of bare TiO2 photocatalyst. The
RGTO achieved 99% degradation of CBZ in 90 min. The improved photocatalytic
activity of RGTO was ascribed to the generation of chemical bonds between
308 A. Kumar et al.

Fig. 12.6  The proposed degradation pathways of DBP in photocatalytic reaction. Reproduced
with permission from [63]. Copyright 2010, Elsevier Publishers

Fig. 12.7 Schematic
presentation of the
proposed mechanism of
CBZ photodegradation.
Reproduced with
permission from [67].
Copyright 2016, Elsevier
Publishers
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 309

graphene oxide (GO) and TiO2. In addition, the macroporous 3D structure of RGOT
aerogel photocatalyst results in mass transportation and abundant surface sites in
CBZ photocatalyst.
Lam and coworkers [69] synthesized the WO3/ZnO nanocomposite using hydro-
thermal route followed by chemical solution process. The successful formation of
the pristine samples and the composites was confirmed. PXRD measurements and
FESEM images for the pristine ZnO nanorods (ZNR) and WO3-ZNR composites
are presented in Fig. 12.8a. The obtained results showed that nanoparticles of mono-
clinic WO3 were homogeneously distributed over ZNR surface. The photocatalytic
activity of pristine ZNR and WO3-ZNR were studied by the decomposition of res-
orcinol (ReOH), which is one of EDC, under visible light irradiations. The experi-
mental results showed that WO3-ZNR photocatalysts exhibited enhanced
photocatalytic performance as compared to ZNR.  The optimized photocatalyst
(WO3 (2 at.%)–ZNR) with 2 at.% WO3 exhibited higher activity towards the degra-
dation of ReOH which was ascribed to the synergic effect of WO3 and ZNR which
prolongs electron-hole separation and their lifetime. The proposed photocatalytic
mechanism for the degradation of ReOH using ZNR-WO3 photocatalyst is pre-
sented in Fig. 12.8b.
The other photocatalyst systems are also reported for the decomposition of EDC
photodegradation. Ding and coworkers [70] have reported the synthesis of Pd/GCN
photocatalysts by using a mechanical mixing-illumination and thermal polymeriza-
tion method. The photocatalytic performance of Pd/GCN catalysts was investigated
through dechlorination of 2-chlorodibenzo-p-dioxin (2-CCD) in the presence of
UV-visible light. It was found that 76% of 2-CDD was converted to dibenzo-p-­
dioxin. Furthermore, kinetics of dechlorination was also studied which shows that
the reaction obeyed pseudo-first-order.
The formation of the Schottky barrier between Pd and GCN benefited the spec-
tral response range and also suppressed the electron-hole pair’s recombination of
GCN (Fig. 12.9). Particularly, 5 wt% Pd modified g-C3N4 exhibited the best dechlo-
rination activity toward 2-CDD.  Our group has also reported a binary

Fig. 12.8 (a) FESEM images of samples (a) ZNR, (b) WO3 (1 at%)–ZNR, (c) WO3 (2 at%)–ZNR,
and (d) WO3 (5 at%)–ZNR. Insert: EDAX of the WO3–ZNR photocatalyst and (b) degradation
mechanism of ReOH using ZNR-WO3 photocatalyst. Reproduced with permission from [69].
Copyright 2012, Elsevier Publishers
310 A. Kumar et al.

Fig. 12.9  Schematic presentation of dechlorination of 2-CDD using Pd/g-C3N4 under UV-visible
light. Reproduced with permission from [70]. Copyright 2018, Elsevier Publishers

nanocomposite (CTCN) comprising of perovskite oxide CaTiO3 and g-C3N4 for the
photocatalytic decomposition of BPA under natural sunlight irradiation [57]. It was
observed that BPA showed very high stability as no degradation was observed dur-
ing the photolysis experiment, whereas CTCN heterojunction degraded about 47%
of BPA in 120 min sunlight irradiation. Petala et al. [71] prepared a novel CuOx/
BiVO4 p/n type heterostructures by coupling the p-type CuO (Cu2O) and n-type
BiVO4 which suppress the electron-hole pair recombination and leads to the effi-
cient generation of superoxide anion and hydroxyl radicals. It was found that 98%
degradation of ethyl paraben (EP) was achieved under sunlight. It was interesting to
note that the presence of bicarbonates results in better photocatalytic decomposition
of EP. Furthermore, the photocatalytic decomposition of EP proceeded through the
hydroxylation and dealkylation processes.

12.4  Ternary and Multi-component Nanostructures

Mukhopadhyay et al. [72] have studied a ternary nanostructure ZnO/TiO2/Au pho-


tocatalyst toward the decomposition of phenanthrene and anthracene. It was found
that 53.7% decomposition of anthracene and 35.6% degradation of phenanthrene
were achieved within 90 min using 10 mg of composite catalyst in the presence of
UV-visible light. ZnO/TiO2/Au ternary nanostructure benefited the visible light
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 311

absorption by combining the surface plasmon resonance effect of Au. In the pres-
ence of UV light, photoexcited electron in the conduction band (CB) of ZnO trans-
fers to the CB of TiO2, whereas holes in valence band (VB), TiO2 moves toward the
VB of ZnO in order to equilibrate the potential difference and result in separation of
photogenerated charge carries. But in case of three component system (ZnO/TiO2/
Au), the photoexcited electron attained extended delocalization from conduction
band of TiO2 to gold nanoparticles (AuNPs). Therefore, well-separated photoex-
cited electron captured by surface adsorbed oxygen molecule (O2) and result in the
formation of O2−• or OH• radical. On the other hand, holes from valence band leads
to the oxidation of H2O/OH− to give OH• radical. These radicals cause the decom-
position of organic pollutants. Ye et al. [73] have reported a novel ternary photocata-
lyst BiVO4-Au@CdS Z-scheme photocatalyst for the decomposition of NP
(nonylphenol) under visible light irradiation. The degradation of NP using
BiVO4-Au@CdS photocatalyst followed pseudo-first-order kinetics. The enhanced
photocatalytic activity was ascribed to the Z-scheme charge transfer which facili-
tates the efficient separation of generated charge carriers and also reserved the
strong redox ability. Furthermore, the photocatalytic mechanism of BiVO4-Au@
CdS Z-scheme and CdS-BiVO4 heterostructure has been presented in Fig. 12.10.
Hung et  al. [74] have investigated the photocatalytic activity of PANI/CNT/TiO2
composite toward the degradation of diethyl phthalate (DEP) under visible light.
The results showed that PANI/CNT/TiO2 composite exhibits high photocatalytic
activity compared to polyaniline and carbon nanotubes (CNT). In general, CNT
result in the suppression of charge carrier recombination, whereas PANI donate
electrons and thus resulting in absorption band edge shiefting of PANI/CNT/TiO2
composite in the visible region.
The quaternary photocatalytic systems are also known for the degradation of
EDC; however, there are only limited reports. Khavar et al. [75] have studied a four-­
component system Fe3O4@rGO@ZnO/Ag-NPs (FGZAg) as a catalyst towards the
photocatalytic decomposition of metformin under both UV and visible light radia-
tion. The results from their findings revealed that complete decomposition and 60%
mineralization of metformin were obtained by using FGZAg photocatalyst within
60 min. Chen et al. [76] synthesized a highly efficient F-Ag-β-CD/TiO2/AC photo-
catalyst with exposed (001) facets by using simple microwave-assisted method. It

Fig. 12.10  Schematic illustration of photocatalytic mechanism of BiVO4-Au@CdS Z-scheme and


CdS-BiVO4 heterostructure. Reproduced with permission from [48]. Copyright 2017, Elsevier
Publishers
312 A. Kumar et al.

was observed that the doping of F into Ag-β-CD/TiO2/AC inhibits the anatase-to-­
rutile phase transformation. The photocatalytic performance of F-Ag-β-CD/TiO2/
AC composite was investigated by degrading naphthalene under visible light. The
enhanced photocatalytic activity was achieved (98.4% degradation) and attributed
to the synergic effect of exposed (001) facets, absorption of visible light, and higher
separation of charge carriers.
In a nutshell, several photocatalysts have been designed and synthesized by the
researchers for the degradation of EDC. The pristine unary photocatalyst generally
showed lower performance owing to the more pronounced recombination of photo-
generated charges and low light absorption. Although, heteroatom doping has been
proven to be beneficial to tackle this problem to some extent along with the light
absorption tailoring. The formation of composites either binary, ternary, or quater-
nary has additional advantages such as choice of selecting suitable semiconductor
components, which offers enhanced light absorption in the solar spectrum and the
proper band positions that allow the efficient charge separation and transfer. The use
of cocatalysts provides active surface sites also and boosts the overall performance
of the working photocatalytic systems. Therefore, the photocatalyst design is very
important to realize the enhanced performance and to develop more efficient materi-
als for EDC removal. Several photocatalysts and their performance towards EDC
removal have been provided in Table 12.1.

12.5  Summary and Perspectives

In this book chapter, the raising concerns of worldwide use of EDC and their harm-
ful effects on the human health have been reviewed. The clean drinking water is the
need of the hour, and therefore the issues originated from the contamination of
water reservoirs must be resolved. The photocatalysis has emerged as a fascinating
process for the removal of EDC from the wastewater. Several photocatalysts have
been used for the photocatalytic degradation of EDC in wastewater, and the reaction
pathways and intermediates generated are well studied in the literature. In compari-
son to the unary and pristine photocatalysts, which are limited by low light absorp-
tion and poor separation of photogenerated charges, the binary and multicomponent
systems have performed well. Although, it is noteworthy that the study of interme-
diates and final TOC values have been reported in some reports. However, most of
the reports do not confirm the complete mineralization of the degraded EDC.  In
addition to the degradation study of individual EDC, the simulated waste water
containing a large number of EDC should also be studied to ensure the effectiveness
of the process. As found in several discussed reports, the complete mineralization of
EDC is still lacking and must be monitored carefully to ensure the complete degra-
dation of the studied samples. The fragments produced during the degradation may
remain undegraded in the final solution which may have higher toxicity than the
parent compounds. Therefore, it is important to study and report the final TOC or
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 313

Table 12.1  Different photocatalysts utilized for EDC degradation


EDC
concentration/
Sl. volume of the Monitoring
no. Photo catalyst Light source EDC solution method Ref.
1. ZnO 8 W medium BPA, BPB, 0.30 mg L−1 UV-vis [7]
pressure Hg lamps DP, BBP, (2000 mL)
(366 nm) MP, EP
2. TiO2 180 W UV LED DP 10 mg L−1 HPLC [77]
(395 nm) (150 mL)
3. TiO2 UV-LED EE2 20 mg L−1 HPLC [78]
(λ = 365 nm) (150 mL)
4. ZnO hierarchical 15 W UV lamps ReOH, MPB, 20 mg L−1 UV-vis and [50]
micro-/ BPA (350 mL) HPLC
nanospheres
5. ZnO and TiO2 150 W Xe lamps BPA, EE2 50–200 μg L−1 UV-vis [79]
6. TiO2 UV lamp MET 0–50 mg L−1 UV-vis and [80]
(250 nm–400 nm) (350 mL) LC-MS
7. TiO2 0.7 kW UV lamp DFC, CPL, 50 mg L−1 UV-vis [81]
(λ = 254 nm) E1, MP
8. C doped ZnO 30 W UV lamp BPA 50 mg L−1 UV-vis and [51]
(λ = 254 nm) (100 mL) LC-MS
9. MWCNT-TiO2 96 W UV lamps DP 1 mg L−1 HPLC [6]
(100 mL)
10. RGO-Ag 250 W Hg vapor Phenol, 100 mg L−1 UV-vis [82]
nanocomposites lamp (visible) and Atrazine, (50 mL)
11 W UV lamp BPA
(UV)
11. WO3 NPs/ZnO 55 W CFL bulbs ReOH 1.8 × 10−4 M UV-vis and [69]
NR (100 mL) HPLC
12. Sn/TiO2 15 W Hg lamp AMOX 20 mg L−1 UV-vis [83]
(λ = 254 nm)
13. BiOBr@SiO2@ 500 W Xe lamp BPA 20 mg L−1 UV-vis and [84]
Fe3O4 (λ > 420 nm) (100 mL) HPLC
14. CD-Fe3O4@TiO2 400 W Hg vapor BPA, DBP 20 ppm LC-MS/ [85]
lamp (17 mW cm−2) MS
15. Ag-WO3/ 450 W Xe arc lamp ATR 20 ppm HPLC [86]
SBA-15 (25 mL)
composites
16. Cu-doped Visible light ATR 100 ppm HPLC and [87]
ZnO/g-C3N4 (400–700 nm) (500 mL) TOC
17. g-C3N4/NiO Visible light CDD 4.59 μmol L−1 HPLC [88]
18. Ag-Fe3O4@ LED bulbs MTF 20 mg L−1 HPLC and [75]
RGO@ZnO (365 nm) and TOC
300 W tungsten Xe
lamp
(continued)
314 A. Kumar et al.

Table 12.1 (continued)
EDC
concentration/
Sl. volume of the Monitoring
no. Photo catalyst Light source EDC solution method Ref.
19. N and La 500 W Xe lamp Naphthalene 20 mg L−1 UV-vis [89]
co-doped TiO2/ (λ < 420 nm) (50 mL)
AC
20. Ag-β-CD/TiO2/ Visible light Naphthalene 30 mg L−1 VU-vis [76]
AC (50 mL)
DP dimethyl Phathalate, BPA bisphenol A, BPB bisphenol B, BBP butyl benzylphthalate, MP
methyl p-hydroxybenzoate, EP 4-hydroxybenzoate, EE2 17β-estradiol, RGO reduced graphene
oxide, ReOH resorcinol, MPB methylparaben, MET metoprolol, NP nanoparticles, DFC diclofe-
nac, CPL chloramphenicol, E1 estrone, MP metoprolol, CD cyclodextrin, DBP dibutyl phthalate,
AMOX amoxicillin trihydrate, ATR atrazine, CDD chlorodibenzo-p-dioxin, MTF metformin

COD values of the degraded samples in addition to the toxicity study of intermedi-
ates observed in HPLC or LCMS and conventional UV–vis spectroscopy monitor-
ing. Some of the EDC in water are in ppm to ppb level; this photocatalytic method
is suitable for complete mineralization upon continuous light irradiation.
Although the photocatalytic decomposition of EDC is fascinating, the results are
limited to the lab scale only. The synthesis of photocatalysts is generally done in few
mg amount in labs, and reaction batches barely exceed 100 mL set up. The scalable
synthesis of the photocatalysts is a big challenge along with the design of large scale
photoreactors. Of course, the large-scale reactor could not be similar to the lab-scale
set up which can encounter the catalyst recovery and loss after each reaction cycle.
The flow reactors could be a good option in which localized catalyst particles
degrade the EDC passing at a certain flow rate in the presence of light irradiation.
Therefore, it is very important to invest more efforts in the design of the reactors to
realize this process at large scale. The collaborative efforts between the chemists
and engineers could pave the way in this direction and future works should be more
focused in this regard.

References

1. Kumar, S., Kumar, A., Kumar, A., & Krishnan, V. (2019). Nanoscale zinc oxide based hetero-
junctions as visible light active photocatalysts for hydrogen energy and environmental reme-
diation. Catalysis Reviews, 62(3), 346–405.
2. World Health Organization. (2017). 2.1 billion people lack safe drinking water at home, more
than twice as many lack safe sanitation. World Health Organization.
3. Hach, C. C., & Buck, M. D. (1977). Method and apparatus for sampling impure water. Google
Patents.
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 315

4. Pi, Y., Li, X., Xia, Q., Wu, J., Li, Y., Xiao, J., & Li, Z. (2018). Adsorptive and photocata-
lytic removal of Persistent Organic Pollutants (POPs) in water by metal-organic frameworks
(MOFs). Chemical Engineering Journal, 337, 351–371.
5. Kumar, A., Kumar, S., Bahuguna, A., Kumar, A., Sharma, V., & Krishnan, V. (2017).
Recyclable, bifunctional composites of perovskite type N-CaTiO 3 and reduced graphene
oxide as an efficient adsorptive photocatalyst for environmental remediation. Materials
Chemistry Frontiers, 1, 2391–2404.
6. Tan, T.  L., Lai, C.  W., Hong, S.  L., & Rashid, S.  A. (2018). New insights into the photo-
catalytic endocrine disruptors dimethyl phathalate esters degradation by UV/MWCNTs-TiO2
nanocomposites. Journal of Photochemistry and Photobiology A: Chemistry, 364, 177–189.
7. Vela, N., Calín, M., Yáñez-Gascón, M. J., Garrido, I., Pérez-Lucas, G., Fenoll, J., & Navarro,
S. (2018). Photocatalytic oxidation of six endocrine disruptor chemicals in wastewater using
ZnO at pilot plant scale under natural sunlight. Environmental Science and Pollution Research,
25, 34995–35007.
8. Vela, N., Calín, M., Yáñez-Gascón, M. J., el Aatik, A., Garrido, I., Pérez-Lucas, G., Fenoll,
J., & Navarro, S. (2019). Removal of pesticides with endocrine disruptor activity in wastewa-
ter effluent by solar heterogeneous photocatalysis using ZnO/Na2 S2 O8. Water, Air, & Soil
Pollution, 230, 134.
9. Canle, M., Pérez, M. I. F., & Santaballa, J. A. (2017). Photocatalyzed degradation/abatement
of endocrine disruptors. Current Opinion in Green and Sustainable Chemistry, 6, 101–138.
10. Spacilova, L., Morozova, M., Masin, P., Maleterova, Y., Kastanek, F., Dytrych, P., Ezechias,
M., Kresinova, Z., & Solcova, O. (2016). Endocrine disruptor degradation by photocatalytic
pilot plant unit. Endocrine, 3, 4613–4620.
11. Vela, N., Calín, M., Yáñez-Gascón, M. J., Garrido, I., Pérez-Lucas, G., Fenoll, J., & Navarro,
S. (2018). Solar reclamation of wastewater effluent polluted with bisphenols, phthalates and
parabens by photocatalytic treatment with TiO2/Na2S2O8 at pilot plant scale. Chemosphere,
212, 95–104.
12. Annamalai, J., & Namasivayam, V. (2015). Endocrine disrupting chemicals in the atmosphere:
Their effects on humans and wildlife. Environment International, 76, 78–97.
13. Darbre, P. D. (2015). What are endocrine disrupters and where are they found? In Endocrine
disruption and human health (pp. 3–26). Elsevier.
14. Darbre, P.  D. (2019). The history of endocrine-disrupting chemicals. Current Opinion in
Endocrine and Metabolic Research, 7, 26–33.
15. Darbre, P. D. (2017). Endocrine disruptors and obesity. Current Obesity Reports, 6, 18–27.
16. Wang, R., Ma, X., Liu, T., Li, Y., Song, L., Tjong, S. C., Cao, L., Wang, W., Qing, Y., & Wang,
Z. (2020). Degradation aspects of endocrine disrupting chemicals: A review on photocatalytic
processes and photocatalysts. Applied Catalysis A: General, 597, 117547.
17. Chu, K. H., Al-Hamadani, Y. A., Park, C. M., Lee, G., Jang, M., Jang, A., Her, N., Son, A., &
Yoon, Y. (2017). Ultrasonic treatment of endocrine disrupting compounds, pharmaceuticals,
and personal care products in water: A review. Chemical Engineering Journal, 327, 629–647.
18. Ullah, A., Hussain, S., Ahmad, W., & Jahanzaib, M. (2020). Development of a decision sup-
port system for the selection of wastewater treatment technologies. Science of The Total
Environment, 731, 139158.
19. Grassi, M., Rizzo, L., & Farina, A. (2013). Endocrine disruptors compounds, pharmaceuti-
cals and personal care products in urban wastewater: Implications for agricultural reuse and
their removal by adsorption process. Environmental Science and Pollution Research, 20,
3616–3628.
20. Kaur, H., Kumar, R., Kumar, A., Krishnan, V., & Koner, R. R. (2019). Trifunctional metal–
organic platform for environmental remediation: Structural features with peripheral hydroxyl
groups facilitate adsorption, degradation and reduction processes. Dalton Transactions, 48,
915–927.
316 A. Kumar et al.

21. Gmurek, M., Olak-Kucharczyk, M., & Ledakowicz, S. (2017). Photochemical decomposi-
tion of endocrine disrupting compounds—A review. Chemical Engineering Journal, 310,
437–456.
22. Hu, X., Zhao, H., Tian, J., Gao, J., Li, Y., & Cui, H. (2017). Synthesis of few-layer MoS2
nanosheets-coated TiO2 nanosheets on graphite fibers for enhanced photocatalytic properties.
Solar Energy Materials and Solar Cells, 172, 108–116.
23. Narváez, J. F., Grant, H., Gil, V. C., Porras, J., Sanchez, J. C. B., Duque, L. F. O., Sossa, R. R.,
& Quintana-Castillo, J. C. (2019). Assessment of endocrine disruptor effects of levonorgestrel
and its photoproducts: Environmental implications of released fractions after their photocata-
lytic removal. Journal of Hazardous Materials, 371, 273–279.
24. Kumar, A., Sharma, V., Kumar, S., Kumar, A., & Krishnan, V. (2018). Towards utilization of
full solar light spectrum using green plasmonic Au–TiOx photocatalyst at ambient conditions.
Surfaces and Interfaces, 11, 98–106.
25. Kumar, A., Kumar, K., & Krishnan, V. (2019). Sunlight driven methanol oxidation by aniso-
tropic plasmonic Au nanostructures supported on amorphous titania: Influence of morphology
on photocatalytic activity. Materials Letters, 245, 45–48.
26. Kumar, A., Reddy, K. L., Kumar, S., Kumar, A., Sharma, V., & Krishnan, V. (2018). Rational
design and development of lanthanide-doped NaYF4@ CdS–Au–RGO as quaternary plas-
monic photocatalysts for harnessing visible–near-infrared Broadband spectrum. ACS Applied
Materials & Interfaces, 10, 15565–15581.
27. Kaur, H., Kumar, A., Koner, R. R., & Krishnan, V. (2020). Metal-organic frameworks for pho-
tocatalytic degradation of pollutants. In Nano-materials as photocatalysts for degradation of
environmental pollutants (pp. 91–126). Elsevier.
28. Reddy, K. L., Kumar, S., Kumar, A., & Krishnan, V. (2019). Wide spectrum photocatalytic
activity in lanthanide-doped upconversion nanophosphors coated with porous TiO2 and Ag-Cu
bimetallic nanoparticles. Journal of Hazardous Materials, 367, 694–705.
29. Kumar, S., Kumar, A., Kumar, A., Balaji, R., & Krishnan, V. (2018). Highly efficient vis-
ible light active 2D-2D nanocomposites of N-ZnO-g-C3N4 for photocatalytic degradation of
diverse industrial pollutants. ChemistrySelect, 3, 1919–1932.
30. Yu, J., Wang, T., & Rtimi, S. (2019). Magnetically separable TiO2/FeOx/POM accelerating
the photocatalytic removal of the emerging endocrine disruptor: 2, 4-dichlorophenol. Applied
Catalysis B: Environmental, 254, 66–75.
31. Shi, M., Luo, L., Dai, J., Xia, L., Long, J., Yang, W., Wang, H., & Shu, L. (2020). The com-
parative study of two kinds of β-Bi 2 O 3/TiO 2 binary composite and their removal of
17ɑ-ethynylestradiol. Environmental Science and Pollution Research, 27(20), 24692–24701.
32. Hak, C.  H., Leong, K.  H., Chin, Y.  H., Saravanan, P., Tan, S.  T., Chong, W.  C., & Sim,
L. C. (2020). Water hyacinth derived carbon quantum dots and gC 3 N 4 composites for sun-
light driven photodegradation of 2, 4-dichlorophenol. SN Applied Sciences, 2, 1–14.
33. Dong, S., Lee, G., Zhou, R., & Wu, J.  J. (2020). Synthesis of g-C3N4/BiVO4 heterojunc-
tion composites for photocatalytic degradation of nonylphenol ethoxylate. Separation and
Purification Technology, 250, 117202.
34. Gao, S., Guo, C., Lv, J., Wang, Q., Zhang, Y., Hou, S., Gao, J., & Xu, J. (2017). A novel 3D
hollow magnetic Fe3O4/BiOI heterojunction with enhanced photocatalytic performance for
bisphenol A degradation. Chemical Engineering Journal, 307, 1055–1065.
35. Meng, A., Zhang, L., Cheng, B., & Yu, J. (2019). Dual cocatalysts in TiO2 photocatalysis.
Advanced Materials, 31, 1807660.
36. Payormhorm, J., & Idem, R. (2020). Synthesis of C-doped TiO2 by sol-microwave method for
photocatalytic conversion of glycerol to value-added chemicals under visible light. Applied
Catalysis A: General, 590, 117362.
37. Javanmardi, S., Nasresfahani, S., & Sheikhi, M. (2019). Facile synthesis of PdO/SnO2/CuO
nanocomposite with enhanced carbon monoxide gas sensing performance at low operating
temperature. Materials Research Bulletin, 118, 110496.
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 317

38. Kumar, A., Navakoteswara Rao, V., Kumar, A., Venkatakrishnan Shankar, M., & Krishnan,
V. (2020). Interplay between mesocrystals of CaTiO3 and edge sulfur atom enriched MoS2
on reduced graphene oxide nanosheets: Enhanced photocatalytic performance under sunlight
irradiation. ChemPhotoChem, 4, 427–444.
39. Rathod, M., Moradeeya, P. G., Haldar, S., & Basha, S. (2018). Nanocellulose/TiO 2 compos-
ites: Preparation, characterization and application in the photocatalytic degradation of a poten-
tial endocrine disruptor, mefenamic acid, in aqueous media. Photochemical & Photobiological
Sciences, 17, 1301–1309.
40. Kotzamanidi, S., Frontistis, Z., Binas, V., Kiriakidis, G., & Mantzavinos, D. (2018). Solar
photocatalytic degradation of propyl paraben in Al-doped TiO2 suspensions. Catalysis Today,
313, 148–154.
41. Al-Hajji, L., Ismail, A. A., Al-Hazza, A., Ahmed, S., Alsaidi, M., Almutawa, F., & Bumajdad,
A. (2020). Impact of calcination of hydrothermally synthesized TiO2 nanowires on their pho-
tocatalytic efficiency. Journal of Molecular Structure, 1200, 127153.
42. Ding, X., An, T., Li, G., Chen, J., Sheng, G., Fu, J., & Zhao, J. (2008). Photocatalytic degra-
dation of dimethyl phthalate ester using novel hydrophobic TiO 2 pillared montmorillonite
photocatalyst. Research on Chemical Intermediates, 34, 67–83.
43. Xu, X.-R., Li, S.-X., Li, X.-Y., Gu, J.-D., Chen, F., Li, X.-Z., & Li, H.-B. (2009). Degradation
of n-butyl benzyl phthalate using TiO2/UV. Journal of Hazardous Materials, 164, 527–532.
44. Muneer, M., Theurich, J., & Bahnemann, D. (2001). Titanium dioxide mediated photocata-
lytic degradation of 1, 2-diethyl phthalate. Journal of Photochemistry and Photobiology A:
Chemistry, 143, 213–219.
45. Su, L., Li, G., Lan, Z., Yu, T., Chu, H., Han, S., & Qin, S. (2020). Preparation of molecularly
imprinted CNT/ZnO and photocatalytic degradation of bisphenol A. Chinese Science Bulletin,
65, 1368–1375.
46. Meenakshi, G., & Sivasamy, A. (2018). Nanorod ZnO/SiC nanocomposite: An efficient cata-
lyst for the degradation of an endocrine disruptor under UV and visible light irradiations.
Journal of Environmental Chemical Engineering, 6, 3757–3769.
47. Jasso-Salcedo, A. B., Hoppe, S., Pla, F., Escobar-Barrios, V. A., Camargo, M., & Meimaroglou,
D. (2017). Modeling and optimization of a photocatalytic process: Degradation of endocrine
disruptor compounds by Ag/ZnO. Chemical Engineering Research and Design, 128, 174–191.
48. Radhika, S., & Thomas, J. (2017). Solar light driven photocatalytic degradation of organic pol-
lutants using ZnO nanorods coupled with photosensitive molecules. Journal of Environmental
Chemical Engineering, 5, 4239–4250.
49. Devaraji, P., Mapa, M., Abdul Hakkeem, H. M., Sudhakar, V., Krishnamoorthy, K., & Gopinath,
C. S. (2017). ZnO–ZnS heterojunctions: A potential candidate for optoelectronics applications
and mineralization of endocrine disruptors in direct sunlight. ACS Omega, 2, 6768–6781.
50. Sin, J.-C., Lam, S.-M., Lee, K.-T., & Mohamed, A. R. (2013). Self-assembly fabrication of
ZnO hierarchical micro/nanospheres for enhanced photocatalytic degradation of endocrine-­
disrupting chemicals. Materials Science in Semiconductor Processing, 16, 1542–1550.
51. Bechambi, O., Sayadi, S., & Najjar, W. (2015). Photocatalytic degradation of bisphenol A in
the presence of C-doped ZnO: Effect of operational parameters and photodegradation mecha-
nism. Journal of Industrial and Engineering Chemistry, 32, 201–210.
52. Zhang, F., Jin, T., Zeng, R., Cui, H., & Song, L. (2014). Cr2O3 nanoparticles modified TiO2
nanotubes for enhancing visible photoelectrochemical performance. Journal of Nanoscience
and Nanotechnology, 14, 7022–7026.
53. You, S., Hu, Y., Liu, X., & Wei, C. (2018). Synergetic removal of Pb (II) and dibutyl phthal-
ate mixed pollutants on Bi2O3-TiO2 composite photocatalyst under visible light. Applied
Catalysis B: Environmental, 232, 288–298.
54. Bahuguna, A., Kumar, A., & Krishnan, V. (2019). Carbon-support-based heterogeneous nano-
catalysts: Synthesis and applications in organic reactions. Asian Journal of Organic Chemistry,
8, 1263–1305.
318 A. Kumar et al.

55. Dhiman, P., Naushad, M., Batoo, K. M., Kumar, A., Sharma, G., Ghfar, A. A., Kumar, G., &
Singh, M. (2017). Nano FexZn1− xO as a tuneable and efficient photocatalyst for solar pow-
ered degradation of bisphenol A from aqueous environment. Journal of Cleaner Production,
165, 1542–1556.
56. Hojamberdiev, M., Czech, B., Göktaş, A. C., Yubuta, K., & Kadirova, Z. C. (2020). SnO2@
ZnS photocatalyst with enhanced photocatalytic activity for the degradation of selected
pharmaceuticals and personal care products in model wastewater. Journal of Alloys and
Compounds, 827, 154339.
57. Kumar, A., Schuerings, C., Kumar, S., Kumar, A., & Krishnan, V. (2018). Perovskite-structured
CaTiO3 coupled with g-C3N4 as a heterojunction photocatalyst for organic pollutant degrada-
tion. Beilstein Journal of Nanotechnology, 9, 671–685.
58. Kumar, A., Naushad, M., Rana, A., Sharma, G., Ghfar, A.  A., Stadler, F.  J., & Khan,

M. R. (2017). ZnSe-WO3 nano-hetero-assembly stacked on Gum ghatti for photo-degradative
removal of Bisphenol A: Symbiose of adsorption and photocatalysis. International Journal of
Biological Macromolecules, 104, 1172–1184.
59. Zhang, L., Yue, X., Liu, J., Feng, J., Zhang, X., Zhang, C., Li, R., & Fan, C. (2020). Facile syn-
thesis of Bi5O7Br/BiOBr 2D/3D heterojunction as efficient visible-light-driven photocatalyst
for pharmaceutical organic degradation. Separation and Purification Technology, 231, 115917.
60. Divya, K., Chandran, A., Reethu, V., & Mathew, S. (2018). Enhanced photocatalytic perfor-
mance of RGO/Ag nanocomposites produced via a facile microwave irradiation for the degra-
dation of Rhodamine B in aqueous solution. Applied Surface Science, 444, 811–818.
61. Hampel, B. R., Kovács, G. B., Czekes, Z., Hernádi, K. R., Danciu, V., Ersen, O., Girleanu, M.,
Focşan, M., Baia, L., & Pap, Z. (2018). Mapping the photocatalytic activity and ecotoxicol-
ogy of Au, Pt/TiO2 composite photocatalysts. ACS Sustainable Chemistry & Engineering, 6,
12993–13006.
62. Li, M., Yu, Z., Hou, Y., Liu, Q., Qian, L., Lian, C., Rao, X., & Yang, X. (2019). Charge trapping
and transfer mechanisms of noble metals and metal oxides deposited Ga2O3 toward typical
contaminant degradation. Chemical Engineering Journal, 370, 1119–1127.
63. Xu, L., Yang, X., Guo, Y., Ma, F., Guo, Y., Yuan, X., & Huo, M. (2010). Simulated sunlight
photodegradation of aqueous phthalate esters catalyzed by the polyoxotungstate/titania nano-
composite. Journal of Hazardous Materials, 178, 1070–1077.
64. Shankar, M., Cheralathan, K., Arabindoo, B., Palanichamy, M., & Murugesan, V. (2004).
Enhanced photocatalytic activity for the destruction of monocrotophos pesticide by TiO2/Hβ.
Journal of Molecular Catalysis A: Chemical, 223, 195–200.
65. Shankar, M.  V., Anandan, S., Venkatachalam, N., Arabindoo, B., & Murugesan, V. (2004).
Novel thin-film reactor for photocatalytic degradation of pesticides in an aqueous solution.
Journal of Chemical Technology and Biotechnology, 79, 1279–1285.
66. Shankar, M., Anandan, S., Venkatachalam, N., Arabindoo, B., & Murugesan, V. (2006). Fine
route for an efficient removal of 2, 4-dichlorophenoxyacetic acid (2, 4-D) by zeolite-supported
TiO2, chemosphere. Chemosphere, 63, 1014–1021.
67. Khavar, A.  H. C., Moussavi, G., Mahjoub, A.  R., Satari, M., & Abdolmaleki, P. (2018).
Synthesis and visible-light photocatalytic activity of In, S-TiO2@ rGO nanocomposite for
degradation and detoxification of pesticide atrazine in water. Chemical Engineering Journal,
345, 300–311.
68. Nawaz, M., Miran, W., Jang, J., & Lee, D.  S. (2017). One-step hydrothermal synthesis of
porous 3D reduced graphene oxide/TiO2 aerogel for carbamazepine photodegradation in
aqueous solution. Applied Catalysis B: Environmental, 203, 85–95.
69. Lam, S.-M., Sin, J.-C., Abdullah, A. Z., & Mohamed, A. R. (2013). ZnO nanorods surface-­
decorated by WO3 nanoparticles for photocatalytic degradation of endocrine disruptors under
a compact fluorescent lamp. Ceramics International, 39, 2343–2352.
70. Ding, J., Long, G., Luo, Y., Sun, R., Chen, M., Li, Y., Zhou, Y., Xu, X., & Zhao, W. (2018).
Photocatalytic reductive dechlorination of 2-chlorodibenzo-p-dioxin by Pd modified g-C3N4
photocatalysts under UV–vis irradiation: Efficacy, kinetics and mechanism. Journal of
Hazardous Materials, 355, 74–81.
12  Nanomaterials for Photocatalytic Decomposition of Endocrine Disruptors in Water 319

71. Petala, A., Bontemps, R., Spartatouille, A., Frontistis, Z., Antonopoulou, M., Konstantinou, I.,
Kondarides, D. I., & Mantzavinos, D. (2017). Solar light-induced degradation of ethyl paraben
with CuOx/BiVO4: Statistical evaluation of operating factors and transformation by-products.
Catalysis Today, 280, 122–131.
72. Mukhopadhyay, S., Maiti, D., Chatterjee, S., Devi, P. S., & Kumar, G. S. (2016). Design and
application of Au decorated ZnO/TiO 2 as a stable photocatalyst for wide spectral coverage.
Physical Chemistry Chemical Physics, 18, 31622–31633.
73. Ye, F., Li, H., Yu, H., Chen, S., & Quan, X. (2018). Constructing BiVO4-Au@ CdS pho-
tocatalyst with energic charge-carrier-separation capacity derived from facet induction and
Z-scheme bridge for degradation of organic pollutants. Applied Catalysis B: Environmental,
227, 258–265.
74. Hung, C.-H., Yuan, C., & Li, H.-W. (2017). Photodegradation of diethyl phthalate with PANi/
CNT/TiO2 immobilized on glass plate irradiated with visible light and simulated sunlight—
Effect of synthesized method and pH. Journal of Hazardous Materials, 322, 243–253.
75. Khavar, A. H. C., Moussavi, G., Mahjoub, A., Yaghmaeian, K., Srivastava, V., Sillanpää, M., &
Satari, M. (2019). Novel magnetic Fe 3 O 4@ rGO@ ZnO onion-like microspheres decorated
with Ag nanoparticles for the efficient photocatalytic oxidation of metformin: Toxicity evalu-
ation and insights into the mechanisms. Catalysis Science & Technology, 9, 5819–5837.
76. Chen, X., Liu, D., Wu, Z., Cravotto, G., Wu, Z., & Ye, B.-C. (2018). Microwave-assisted rapid
synthesis of Ag-β-cyclodextrin/TiO2/AC with exposed {001} facets for highly efficient naph-
thalene degradation under visible light. Catalysis Communications, 104, 96–100.
77. Ku, Y., Shiu, S.-J., & Wu, H.-C. (2017). Decomposition of dimethyl phthalate in aqueous solu-
tion by UV–LED/TiO2 process under periodic illumination. Journal of Photochemistry and
Photobiology A: Chemistry, 332, 299–305.
78. Arlos, M.  J., Liang, R., Hatat-Fraile, M.  M., Bragg, L.  M., Zhou, N.  Y., Servos, M.  R., &
Andrews, S.  A. (2016). Photocatalytic decomposition of selected estrogens and their estro-
genic activity by UV-LED irradiated TiO2 immobilized on porous titanium sheets via thermal-­
chemical oxidation. Journal of Hazardous Materials, 318, 541–550.
79. Zacharakis, A., Chatzisymeon, E., Binas, V., Frontistis, Z., Venieri, D., & Mantzavinos,
D. (2013). Solar photocatalytic degradation of bisphenol A on immobilized ZnO or TiO2.
International Journal of Photoenergy, 2013, 570587.
80. Romero, V., Marco, P., Giménez, J., & Esplugas, S. (2013). Adsorption and photocatalytic
decomposition of the-blocker metoprolol in aqueous titanium dioxide suspensions: Kinetics,
intermediates, and degradation pathways. International Journal of Photoenergy, 2013, 138918.
81. Czech, B., & Rubinowska, K. (2013). TiO 2-assisted photocatalytic degradation of diclofenac,
metoprolol, estrone and chloramphenicol as endocrine disruptors in water. Adsorption, 19,
619–630.
82. Bhunia, S. K., & Jana, N. R. (2014). Reduced graphene oxide-silver nanoparticle composite
as visible light photocatalyst for degradation of colorless endocrine disruptors. ACS Applied
Materials & Interfaces, 6, 20085–20092.
83. Mohammadi, R., Massoumi, B., & Rabani, M. (2012). Photocatalytic decomposition of amox-
icillin trihydrate antibiotic in aqueous solutions under UV irradiation using Sn/TiO2 nanopar-
ticles. International Journal of Photoenergy, 2012, 514856.
84. Zhang, L., Wang, W., Sun, S., Sun, Y., Gao, E., & Zhang, Z. (2014). Elimination of BPA
endocrine disruptor by magnetic BiOBr@ SiO2@ Fe3O4 photocatalyst. Applied Catalysis B:
Environmental, 148, 164–169.
85. Chalasani, R., & Vasudevan, S. (2013). Cyclodextrin-functionalized Fe3O4@ TiO2: Reusable,
magnetic nanoparticles for photocatalytic degradation of endocrine-disrupting chemicals in
water supplies. ACS Nano, 7, 4093–4104.
86. Gondal, M., Suliman, M., Dastageer, M., Chuah, G.-K., Basheer, C., Yang, D., & Suwaiyan,
A. (2016). Visible light photocatalytic degradation of herbicide (Atrazine) using surface plas-
mon resonance induced in mesoporous Ag-WO3/SBA-15 composite. Journal of Molecular
Catalysis A: Chemical, 425, 208–216.
320 A. Kumar et al.

87. Truc, N. T. T., Duc, D. S., Van Thuan, D., Al Tahtamouni, T., Pham, T.-D., Hanh, N. T., Tran,
D. T., Nguyen, M. V., Dang, N. M., & Le Chi, N. T. P. (2019). The advanced photocatalytic
degradation of atrazine by direct Z-scheme Cu doped ZnO/g-C3N4. Applied Surface Science,
489, 875–882.
88. Ding, J., Lu, S., Shen, L., Yan, R., Zhang, Y., & Zhang, H. (2020). Enhanced photocatalytic
reduction for the dechlorination of 2-chlorodibenzo-p-dioxin by high-performance g-C3N4/
NiO heterojunction composites under ultraviolet-visible light illumination. Journal of
Hazardous Materials, 384, 121255.
89. Liu, D., Wu, Z., Tian, F., Ye, B.-C., & Tong, Y. (2016). Synthesis of N and La co-doped TiO2/
AC photocatalyst by microwave irradiation for the photocatalytic degradation of naphthalene.
Journal of Alloys and Compounds, 676, 489–498.
Chapter 13
Carbonaceous Nanomaterials
for Environmental Remediation

Natarajan Sasirekha and Yu-Wen Chen

13.1  Introduction

Global environmental issues like depletion of fossil fuel resources, climate change,
environmental degradation, and pollution are some of the consequences of industri-
alization and urbanization, which need a systematic approach using suitable reme-
diation technologies. Environmental remediation by integrating sustainability has
been an immense scientific challenge in the recent years in order to balance the
effects and benefits of the environment, economy, and society. The choice of reme-
dial technology to mitigate environmental pollution such as water decontamination,
soil reclamation, and air purification depends mainly upon the contaminants of con-
cern and their allowable concentrations in a specific site and the efficacy of the
technology to reduce their concentrations to allowable limits in a cost-effective
manner. Also, the ease of implementation of the remediation technology at the pol-
luted site should also be considered. The promising approach is to adapt innovative
and recent technologies to find solutions to emerging environmental issues.
Nanotechnology is a fascinating field of interest because nanomaterials exhibit
unique physical and chemical properties that can be tuned to address some of the
potential problems in different fields of interest. Significant breakthroughs have
already occurred in various fields such as energy, biomedicine, environment, cataly-
sis, engineering, sensing, and electronics due to their high surface area, high reactiv-
ity, easy dispersibility, and rapid diffusion. However, the potential of nanotechnology
in combating the burning environmental issues has not been explored thoroughly
compared with other fields of interest. Researchers are still working hard to improve

N. Sasirekha (*)
CAS in Crystallography and Biophysics, University of Madras, Guindy Campus,
Chennai, Tamil Nadu, India
Y.-W. Chen
Department of Chemical and Materials Engineering, National Central University,
Chung-Li, Taiwan

© Springer Nature Switzerland AG 2021 321


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_13
322 N. Sasirekha and Y.-W. Chen

Fig. 13.1  Overview of the chapter

the efficiency of nanomaterials by modifying their physico-chemical properties


using different preparation techniques. Metal-free carbonaceous nanomaterials
have attracted interest for applications such as electronic applications, environmen-
tal remediation, biological imaging, and sensing, after discovering their extraordi-
nary and tunable mechanical, electrical, thermal, and optical properties.
In this chapter, we have presented a comparative survey of the environmental
applications of carbonaceous nanomaterials such as carbon nanotubes, fullerenes,
graphene and their composites, and nanoporous carbon materials (Fig. 13.1). The
recent literature reports on carbonaceous nanomaterials based on their structure and
types toward decontamination of water and soil polluted by emerging contaminants
like antibiotics, pesticides, dyes, heavy metals, etc., and air polluted by toxic inor-
ganic gases and volatile organic compounds (VOCs) are discussed. Furthermore,
various techniques available for the treatment of pollutants using carbon nanomate-
rials are also considered.

13.2  Carbonaceous Nanomaterials

13.2.1  Types of Carbonaceous Nanomaterials

Carbon is a fascinating element that forms the backbone of life on earth. It is known
to bond between them using the hybridization of their 2s-2p orbitals, resulting in
various carbon allotropes having sp, sp2, and sp3 hybridization. Its natural allotropes
are diamond and graphite (Fig. 13.2). The carbon atoms in the case of diamond and
13  Carbonaceous Nanomaterials for Environmental Remediation 323

Fig. 13.2 Carbon
allotropes

Synthetic
Carbon
Allotropes

graphite have sp3 hybridized carbon network and stacked layers of sp2 hybridized
carbon atoms, respectively. After the discovery of buckminsterfullerene in 1985 by
Curl, Kroto, and Smalley, more attention was given toward the synthesis of syn-
thetic nanocarbon allotropes, which paved way for the discovery of carbon nano-
horns [1], carbon nano-onions (multi-shell fullerenes), carbon nanotubes [2],
graphene [3], nanodiamonds, carbon dots [4], graphene (quantum) dots, etc. These
carbon allotropes are classified into zero-dimensional (0D), one-dimensional (1D),
and two-dimensional (2D) based on dimensionality (Fig. 13.2). Some of the synthe-
sized carbonaceous nanomaterials include (1) 0D fullerenes (C60), fullerene-struc-
tures (C70, C76, C84, C90), carbon nano-­onions, carbon dots, graphene dots, and
nanodiamond; (2) 1D CNTs, carbon nanohorns and nanofibers; (3) 2D graphene,
graphene nanoribbons; and (4) ordered mesoporous carbon. Graphene and diamond
structures are the origin of most of the carbon allotropes. Apart from the abovemen-
tioned carbon allotropes, there are several carbon allotropes predicted using theo-
retical approaches.
Graphene, which is a flat single layer of graphite, can be considered as the mother
for most of the synthetic carbon allotropes. Graphene can be rolled, wrapped, and
stacked to form CNTs, fullerenes, and graphite, respectively. The unique feature of
graphene lies in its pure sp2 hybridized state of bonded carbon atoms, exclusively
connected with covalent bonds that result in strong mechanical stability (C–C bond
length 1.42 Å). The other forms of carbon allotropes have various degrees of rehy-
bridization in order to accommodate their shape and structure. Most of them contain
some sp3 carbon atoms along the edges or defect sites apart from the primary sp2
carbon atoms. However, nanodiamond has a mixture of amorphous and graphitic
regions, and hence, it contains both sp3 and sp2 carbon atoms in various ratios. All
these carbon allotropes possess unique properties different from graphene because
of their distinctive shapes. Apart from the most stable 2D carbon form, graphene,
different 2D carbon allotropes such as graphynes, graphdiyne, graphenylene, biphe-
nylene, ψ-graphene, etc. are also proposed using computational methods.
324 N. Sasirekha and Y.-W. Chen

CNTs are produced by rolling up of the one-atom-thick graphene into a cylinder.


It can be classified based on the number of graphenic layers in the tube wall into
single-walled carbon nanotubes (SWCNTs), double-walled carbon nanotubes, and
multi-walled carbon nanotubes (MWCNTs). By changing the orientation of the
honeycomb-like carbon plane in a specific direction, i.e., chirality, it can form zig-
zag, armchair, and chiral nanotubes. The latest addition to carbonaceous 1-D nano-
material is phenine nanotubes (pNT), an analog of CNT that consists of
1,3,5-trisubstituted benzene (phenine) units instead of trigonal sp2 carbon atoms in
CNT with periodic six atoms vacancy defects. The cylindrical graphitic sheet con-
sists of 240 sp2 carbon atoms [5]. The pores open up new avenues for entrapment of
various molecules. The physico-chemical properties of the fascinating pNT are yet
to be explored. Carbon nanohorns (nanocones) are nanotubes with conical tips, con-
structed from sp2 hybridized graphene. Carbon nanofiber is a quasi-1D carbon that
is non-tubular and non-continuous in nature with graphitic layers.
Fullerenes are nothing but a closed hollow cage formed by wrapping graphene
with sp2 hybridized carbon atoms that are arranged in the form of pentagons and
hexagons. It is also called as buckyballs. C60 is the smallest and most widely studied
0D carbon allotrope and consists of 12 pentagons and 20 hexagons to form trun-
cated icosahedron. The arrangement of concentric graphenic shells, similar to onion
layers in onion, is called as carbon nano-onions or multi-shelled fullerenes. The
emerging quasi-spherical carbon-based nanodots are carbon quantum dots
(2–10 nm) and graphene quantum dots (1–10 nm). Carbon dots are formed by a dif-
ferent combination of graphitic and turbostratic carbon and are mostly amorphous
in nature with sp3 hybridized carbon atoms. In contrast, graphene quantum dots are
crystalline in nature with sp2 hybridized carbon atoms. Nanodiamonds consist of sp3
amorphous core and sp2 graphene shell with a truncated octahedral structure.

13.2.2  Properties

Graphene and its derivatives exhibit unusual electronic and spintronic properties
apart from the high theoretical specific surface area (2630 m2 g−1), good mechanical
(Young’s modulus of 1.0 TPa) and thermal stability, ambipolar field effect, high
transparency and high intrinsic electrical conductivity (106 S cm−1), charge mobility
(200,000 cm2 V−1 s−1), and thermal conductivity (~5000 W m−1 K−1). It is a good
conductor as it has a zero bandgap with low optical absorbance of 2.3%. It is quite
stable and expresses inertness to chemical reactions. CNT has a variable bandgap
between zero and 2 eV based on the structure. Hence, its properties can vary from
metallic to semiconducting behavior. The thermal conductivity of SWCNT
(6000 W m−1 K−1) is higher than MWCNT (2000 W m−1 K−1). The electrical prop-
erty of armchair CNT is metallic, whereas the chiral and zigzag are semiconductors
in nature. The properties of different allotropes and their modifications have been
studied and discussed extensively in several review papers [6–9].
13  Carbonaceous Nanomaterials for Environmental Remediation 325

13.2.3  Synthesis

Graphene is synthesized by top-down or bottom-up approach from graphite or from


solution atoms, respectively (Fig. 13.3). One of the oldest and simple methods to get
graphene sheets from graphite is by mechanical cleavage of graphite using a scotch
tape method. However, it would be difficult to get a single graphene sheet by this
method. Chemical vapor deposition (CVD) on metal surfaces is found to be a prom-
ising synthesis method for graphene. Other synthesis methods include chemical
coupling reactions or exfoliation of graphite powder via solution oxidation, sonica-
tion, intercalation, and ball milling. Even though graphene exhibits excellent
physico-chemical properties, pristine graphene has poor solubility and fusibility. It
has the capacity to stack together through π–π and hydrophobic interactions to form
wrinkles. Hence, several methods were developed to improve the surface and elec-
tronic structure of graphene by chemical functionalization in order to expand and
scale up the applications of graphene to industrial scale in various fields of research.
The bandgap of graphene can be altered by doping, intercalation, and stripping
methods.

Fig. 13.3  Techniques for the synthesis of graphene and CNT


326 N. Sasirekha and Y.-W. Chen

The functionalization possibilities in graphene structure are through (1) edge


sites of graphene with dangling bonds, (2) basal plane of strong covalent bonding
with highly delocalized π electrons, (3) non-covalent adsorption on basal plane, (4)
asymmetrical functionalization of the basal plane, and (5) self-assembling of func-
tionalized graphene sheets. The limitations of pristine graphene can be overcome by
adopting suitable chemical functionalization. The covalent bond functionalization
of graphene with functional groups containing oxygen forms graphene oxide (GO),
which contains a combination of sp2 and sp3 carbon atoms.
GO is synthesized by Hummer’s method from graphite oxidation using NaNO3
and KMnO4 dissolved in concentrated H2SO4. It contains hydroxyl and epoxy
groups at the basal plane and a carboxyl group at its edge. The reduced form of GO
is called reduced graphene oxide (rGO). The functionalization of GO can take place
at epoxy groups through ring opening reactions or at carboxylic acid groups through
the addition of nucleophilic species or non-covalent functionalization through π–π
interactions.
CNTs are produced by different techniques such as arc discharge, laser evapora-
tion, catalytic pyrolysis, flat method, template method, and flame method [9]
(Fig. 13.3). CVD, matrix method, and spray method are some of the catalytic pyrol-
ysis methods. Similar to graphene, CNTs also need functionalization to improve
solubility and other physico-chemical properties required based on the desired
applications. Functionalization can be carried out by attaching functional groups or
polymer chains on CNTs. It can be classified into (1) endohedral functionalization,
where the empty inner cavity of CNT is filled with different molecules or nanopar-
ticles, and (2) exohedral functionalization, where the outer surface of CNT is grafted
with molecules. Fullerenes can also be used to fill the inner cavity of CNT. Based
on the approach of functionalization, viz., defect functionalization, covalent func-
tionalization (end group, defect group, side-wall functionalization), and non-­
covalent interaction (van der Waals force), different sites are available on CNTs.
The common terminal group is –COOH and less common are –OH, –H, and
=O. Apart from free-standing graphene and CNTs for various applications, they can
also be modified by grafting with suitable organic compounds or dispersed on a
suitable support or converted into a core-shell morphology with metal oxides/metals
or multi-modified to form hybrid materials (Fig. 13.4).
In the case of supported CNTs, the preparation of CNTs usually ends up with
different morphologies such as aggregated CNTs, vertically aligned CNTs
(VACNTs), and horizontally aligned CNTs (HACNTs). However, both aggregated
CNTs and VACNTs exhibit structural defects. HACNTs can grow up to a few cen-
timeters or decimeters in length with low defect densities. It can also be classified
into SWCNT and MWCNT. Among the synthesis of SWCNT and MWCNT, it is
difficult to synthesize SWCNT in bulk quantities.
The close structural analog of carbon nanohorn is SWCNT. The main challenge
in producing defect-free CNT lies in the complete removal of metal particles during
the production of CNT. It needs strong acid treatment that may end up in defect
formation and loss of carbon material. However, the synthesis of CNH is an easy
and simple method without the necessity to add metal particles. Also, it can be
13  Carbonaceous Nanomaterials for Environmental Remediation 327

Fig. 13.4 Functionalization
of carbonaceous
nanomaterials. (a) Grafted,
(b) Supported/Coated,
(c) Core-shell, (d) Multi-
modification [10]

Fig. 13.5  Techniques for


the synthesis of
carbon dots

produced at room temperature. Hence, CNH could be an alternative to some of the


applications of CNTs.
Fullerenes are produced from graphite by vaporization using arc and plasma
discharge, laser irradiation, naphthalene pyrolysis, and hydrocarbon combustion
[11, 12]. Carbon quantum dots (CDs) are prepared from top-down and bottom-up
approaches. The top-down approach includes fragmentation of carbon source by arc
discharge, laser ablation, exfoliation by chemical and electrochemical oxidation,
ultrasonic exfoliation, whereas the bottom-up approach includes carbonization of
carbon precursors by solvothermal, pyrolysis, and microwave-assisted method
(Fig. 13.5). Doping is used to tune the optical properties of CDs by metal atoms and
heteroatoms. Heteroatom doping can be single heteroatom doping (N-doped CDs,
S-doped CDs, B-doped CDs) and co-doping multiplex heteroatoms (N,P-co-CDs,
N,S-co-CDs, Mg,N-co-doped, Co-N-co-doped) [9]. CDs contain hydroxyl, car-
bonyl, carboxyl, epoxy, and ether functional groups and are adaptable for function-
alization using amine, phosphorous, sulfur, and boron-containing heteroatoms.
328 N. Sasirekha and Y.-W. Chen

Nanodiamonds are synthesized by detonation techniques, laser ablation, high-­


energy ball-milling of high-pressure, high temperature, CVD, etc. Detonation is the
preferred technique, where an explosive mixture of trinitrotoluene and hexogen
(1,3,5-trinitro-1,3,5-triazinane) is detonated. The nanodiamond tends to aggregate
and also, the product contains incombustible impurities along with the graphitic
nanodiamond, which needs to be purified before any application [8].
In general, the physico-chemical properties of carbonaceous nanomaterials can
be tuned by surface doping, interior doping, and introduction of functional groups.
Pristine graphene, CNTs, and CNOs have poor solubility in water and organic sol-
vents that need to be overcome by functionalization and suitable modifications to
decrease surface hydrophobicity. Functionalization can be carried out by pure
chemical methods (chemical oxidation, deposition) or by extended chemical meth-
ods such as electrochemical, microemulsion, sol-gel, and hydrothermal methods.
Activated carbon is a form of porous carbon material formed by carbonization of
an organic precursor followed by physical activation or chemical activation. In gen-
eral, it comprises mainly of micropores along with a negligible amount of meso-
pores and micropores. The porous structure facilitates the high surface area of
activated carbon. It is a rather simple and cost-effective process that attracts indus-
trial applications. However, the non-uniform porous structure and dominant micro-
porous structure limit certain applications of activated carbon. Ordered mesoporous
carbon has been used to overcome the demerits of activated carbon due to periodic
pore arrangement, high surface area, and large pore volume. The template method
is a versatile approach used for the synthesis of ordered mesoporous carbon through
either hard template or soft template (Fig.  13.6). In hard template, an inorganic
mesoporous oxide (SBA-15) has to be taken as a template or mold, and carbon pre-
cursor (sucrose) has to be introduced. Followed by carbonization, the template will
be removed to form ordered mesoporous carbon. In the case of the soft template
method, an assembly of amphiphilic molecules arranged to form a hexagonal
structure, followed by surfactant removal and calcination. The porosity depends
upon the template method as well as the choice of precursors.

Fig. 13.6  Synthesis of ordered mesoporous carbon materials by template method


13  Carbonaceous Nanomaterials for Environmental Remediation 329

13.3  Carbonaceous Nanomaterials


for Water Decontamination

Water, the elixir of life, is the most valuable natural resource that has to be preserved
from various pollutants that are discharged intentionally or unintentionally into
water bodies. Due to population explosion, urbanization, and industrialization,
water bodies get polluted rapidly because of marine dumping, sewage and wastewa-
ter, industrial effluents, agricultural run-off, radioactive waste, underground storage,
leakage, household waste, oil pollution, atmospheric deposition, etc. It is the need
of the hour to improve the quality of the existing available water resources, both
surface and groundwater. In general, water pollutants can be classified into organic
wastes and inorganic wastes, based on the nature of pollutants. However, it is quite
sensible to demarcate the pollutants into conventional pollutants such as metals and
metalloids, and emerging water pollutants. Emerging contaminants, which need
immediate attention include pharmaceuticals, personal care products, endocrine
disruptors, artificial sweetener, and surfactants. Inorganic metals and metalloids are
essential to the growth of plants and human beings in trace amounts; however, the
concentrations above the permissible standard limits pose serious effects on the
ecosystem and human health.
In general, wastewater treatment includes physical, biological, and chemical
treatment, where physical contaminants, organic pollutants, and microbes are
removed. However, the removal of metals and emerging contaminants need special
treatment. Biological treatment is effective only on biodegradable contaminants,
whereas physical and chemical methods result in a large quantity of sludge. The
advanced oxidation process (AOP) is a promising technology for the removal of
non-biodegradable and toxic substances in water and wastewater, by the generation
of highly reactive oxygen species in the reaction medium. It is considered to be a
safe and environment-friendly method without generating any secondary pollutants.
Some of the common AOP methods are direct photolysis, homogeneous photolysis,
heterogeneous photolysis, photo-Fenton, ozonation, and photocatalysis.
It is mandatory to adopt emerging technology to effectively remove the conven-
tional and emerging pollutants, along with a method for water recycling. Due to
high surface-to-volume ratio and unique properties, nanomaterials can completely
degrade or convert toxic substances into a less hazardous one. Nanomaterials can
act in the form of adsorbents, catalysts, and membranes for water decontamination
processes such as adsorption, photocatalysis, membrane process, desalination, dis-
infection, and monitoring. Carbonaceous nanomaterials have the advantages of high
surface area, nanodimensions, and excellent physico-chemical properties. They can
be modified by functionalization and can also be used to form nanocomposites to
exploit the advantage of synergistic interactions. The chemical inertness and bio-
compatibility of carbonaceous nanomaterials make them a suitable candidate for
wastewater treatment.
330 N. Sasirekha and Y.-W. Chen

13.3.1  Adsorption

Adsorption is a surface phenomenon, where pollutants interact with the surface of


carbon materials by electrostatic attraction, weak interactions (van der Waals forces)
or chemical bonding, based on the nature of adsorbent and adsorbate, to change
from liquid/gaseous pollutant to solid state. During adsorption, the surface energy
tends to decrease because of reduction in the number of exposed surface atoms.
Adsorption is considered as a preferred method for environmental remediation due
to its low cost, simple method, trace level pollutant removal, efficient, suitable for
continuous and batch process, easy availability of materials as adsorbents, regenera-
tion and reuse of adsorbents, and absence of by-products during the process.
Nanoadsorbents are alternative to conventional adsorbents owing to their large sur-
face area, small size, tunable surface properties, high adsorptive capacity, high
affinity toward adsorbates, small internal diffusion resistance, negligible intraparti-
cle diffusion distance, magnetic and optical properties, and catalytic properties.
The physico-chemical properties of carbon nanomaterials such as graphene,
CNT, fullerene, etc. can be tuned by surface modifications to improve dispersion in
various medium, adsorption capacity, affinity, and selectivity towards adsorbate.
The surface modifications can be carried out by chemical reactions or with the aid
of organic supports. It can also be combined with transition metal oxides, nanoma-
terials to form nanocomposite or nanohybrids. The adsorption mechanism of carbon
nanomaterials depends upon the nature of pollutants. Organic pollutants tend to
interact through physical adsorption, electrostatic interaction, cation–π bonding,
complexation (inner sphere, surface), ion-exchange and precipitation, whereas inor-
ganic metals and metalloids bind due to hydrophobic effect of adsorbent, electron
donor acceptor (EDA) interaction, electrostatic interaction, hydrogen bonding, par-
titioning, and pore filling (Fig. 13.7).
For example, GO contains oxygen comprising functional groups such as epoxy,
hydroxyl, carbonyl, and carboxyl groups that can enhance the hydrophilicity of
GO. The extreme hydrophilicity of GO makes it a suitable candidate for the removal
of organic and inorganic pollutants from wastewater. Figure 13.8 depicts the inter-
actions of organic pollutants and inorganic contaminants with graphene-based
nanoadsorbents through cation attraction, anion attraction, hydrophobic interac-
tions, ion exchange, hydrogen bonding, and π–π bonding (EDA interactions).
The specific interaction for a particular adsorbent and adsorbate is also decided
based on the physical and chemical conditions of the medium of interaction such as
pH, temperature, concentration, etc. In order to understand the relationship between
surface properties of adsorbent and concentration of adsorbate under specific condi-
tions, several adsorption isotherm models such as Langmuir, Freundlich, Dubinin–
Radushkevich, Temkin, and Linear models were applied. The kinetics studies and
interpreting the mechanism of adsorption processes are inevitable in optimizing the
reaction conditions and parameters to scale up to industrial level.
13  Carbonaceous Nanomaterials for Environmental Remediation 331

Fig. 13.7  Sorption mechanism for the interaction of organic compounds and inorganic metals
with carbonaceous nanomaterials

Fig. 13.8  Schematic representation of aqueous removal of inorganic and organic contaminants by
graphene-based nanoadsorbents [13]

Degradation of Organic Pollutants

Nowadays, more attention is given toward the elimination of organic pollutants such
as persistent organic pollutants (polychlorinated biphenyls, dioxins, DDT, HCH),
endocrine disruptive chemicals (pesticides, dyes), and pharmaceuticals, due to its
high bioaccumulation, biomagnification, and carcinogenic characteristics. Several
adsorbents have been used for the degradation of organic pollutants; however, the
interesting properties of carbonaceous nanomaterials attracted the researchers to
focus on graphene-based materials, CNTs, fullerene, and nanoporous carbon mate-
rials for adsorption.
Both graphene and CNT can form physical interactions with organic compounds
through van der Waals forces, electrostatic forces, or π-bonds, whereas chemical
bonding is possible only after functionalization on the basal surface. The dominant
332 N. Sasirekha and Y.-W. Chen

interaction is based on the surface properties of the adsorbent and the organic pol-
lutants. Pure graphene is hydrophobic and facilitates the adsorption of hydrocar-
bons. The chemical functionalization creates hydrophilic sites based on the extent
of functionalization that eases the adsorption of organic compounds containing both
hydrophobic and hydrophilic groups. The main advantage of graphene as an adsor-
bent is that it can provide both sides of the sheet for adsorption. In the case of CNTs,
the external surface and hollow core are used for adsorption. The agglomerates of
graphene and CNTs can provide a porous structure containing micropores, meso-
pores, and macropores that can enhance their adsorption capacity [14]. The wedge-­
shaped or stilt-shaped pores that are found in the aggregates are flexible compared
with the fixed pores in activated carbon. The large surface area improves π–π EDA
interactions and adsorption capacity, but an increase in the curvature of CNTs
decreases adsorption capacity. The application of graphene and GO-based compos-
ites for the removal of organic pollutants have been well-reviewed in the literature
[15, 16].

Degradation of Antibiotics

The application of antibiotics in healthcare and veterinary medicine is inevitable. It


has been observed that antibiotics and their partially degraded metabolites were
released into the environment in a considerable amount. Antibiotics are classified
based on pharmacological properties as β-lactams, sulfonamides, monobactams,
carbapenems, aminoglycosides, glycopeptides, lincomycin, macrolides, polypep-
tides, polyenes, rifamycin, tetracyclines, chloramphenicol, quinolones, and fluoro-
quinolones. Due to the incomplete metabolism of these antibiotics in human and
animals resulted in residual antibiotics in surface water and wastewater stream that
leads to the development of antimicrobial-resistant organisms by genetic mutations.
The conventional treatment methods are not effective in removing antibiotics and
their metabolites, and hence, it is necessary to find an effective technique to remove
this emerging and pseudo-persistent organic pollutant. Different techniques are
available for the removal of antibiotics such as adsorption, photodegradation, oxida-
tion, and biodegradation. Adsorption is a simple, cost-effective, and efficient
method. Scientists have attempted nanoclays, chitosan, metal oxides, activated car-
bon, and other nanoporous materials as adsorbents for the removal of antibiotics.
The development of novel adsorbents is still in demand for the improvement of
efficient and cost-effective treatment technology.
One of the widely used antibiotics for bacterial infections in humans is fluoroqui-
nolones. Ciprofloxacin, which is an efficient, cost-effective, and essential medicine,
is a fluoroquinolone antibiotic used only in humans. It is noted that fluoroquino-
lones are excreted from the human body without undergoing any change or partially
metabolized. The ciprofloxacin residue is detected in the surface and drinking water.
Being a highly dissolvable antibiotic at various pH ranges, the degradation of cipro-
floxacin is a challenge. The other common antibiotics that are found in water are
sulfonamides (sulpha drugs) and tetracyclines, which are used for treating
13  Carbonaceous Nanomaterials for Environmental Remediation 333

Fig. 13.9  Degradation of antibiotics by carbon nanomaterials

infectious bacterial diseases in human and veterinary medicine. This chapter focuses
on the degradation of fluoroquinolones, sulfonamides, and tetracyclines using car-
bon nanomaterials, with special emphasis on ciprofloxacin, sulfamethoxazole, and
chlorotetracycline (Fig. 13.9).
Ciprofloxacin is highly stable in water that complicates the degradation effi-
ciency. Literature reports on hybrid carbonaceous nanomaterials have found to be
promising toward the adsorption of pollutants. Zhuang et al. studied the effect of
long TiO2 nanotube supported on reduced graphene oxide (rGO–TON) hydrogel for
the adsorption of ciprofloxacin. GO has good adsorption capacity but poor regen-
eration capacity. The presence of TON reduces the agglomeration of graphene
sheets and increases the specific surface area (138.2  m2g−1). The 3D structure of
rGO–TON facilitates good adsorption and regeneration capacity than rGO and
rGO–P25. The adsorption studies fit well with Langmuir model and the maximum
adsorption capacities of rGO, rGO–TON, and rGO–P25 are 178.6, 181.8, and
108.7  mg.g−1, respectively [17]. GO has also been combined with metal organic
frameworks (MOF) to form composites, in order to exploit the advantage of MOF
such as high surface area, crystalline architecture, porous structure, and large active
sites for the adsorption of ciprofloxacin. Manganese-loaded Prussian blue analogs
(Mn-PBA) on GO were prepared by the solvothermal method. Khan et al. reported
that the adsorption capacity of Mn-PBA/GO outperforms the adsorption capacity of
already reported adsorbents [18].
Ordered mesoporous carbon (OMC) fabricated using SBA-15 and pitch resin
and modified with chemical agents such as phosphoric acid, zinc chloride, or ammo-
nium hydroxide was used for adsorption of ciprofloxacin [19]. In general, cipro-
floxacin can exist as cation, zwitterion, or anion form based on the pH of the reaction
mixture. The chemical modification of hydrophobic OMC increased the number of
acidic or basic groups on the carbon surface, which alters the adsorption capacity of
ciprofloxacin. The ciprofloxacin contains a fluorine group that provides a strong
electron-withdrawing ability. The highly insoluble ciprofloxacin can adsorb easily
on the hydrophobic surface of OMC, and hence hydrophobic surface interactions
should be a dominant mechanism, apart from π–π electron donor–acceptor and
hydrogen bonding interactions. The modified OMC showed better adsorption with
334 N. Sasirekha and Y.-W. Chen

a neutral or zwitter ionic form of ciprofloxacin than unmodified OMC. The adsorp-


tion follows the Langmuir adsorption model and pseudo-second-order kinetics.
Intraparticle diffusion and boundary layer control also affect the adsorption process.
Veclani and Melchior studied the neutral and zwitter ionic forms of ciprofloxacin
adsorbed on SWCNT by molecular dynamics simulations [20]. According to the
results, the inner wall of SWCNT has more affinity toward ciprofloxacin and, hence,
maximum adsorption at the inner wall. The adsorption of ciprofloxacin oriented
parallel to the surface of SWCNT was mainly through π–π interactions.
The wastewater is usually mixed with different pollutants to be removed.
Recently, Ma et al. investigated the adsorption capacity of graphene hydrogels (GH)
toward the adsorption of ciprofloxacin (CIP) in the presence of heavy metal, Cu(II).
At low concentrations, the presence of Cu(II) weakly enhances the adsorption of
ciprofloxacin by bridging effect, whereas, at high concentrations, there exists a
competition between ciprofloxacin and Cu(II) for adsorption sites, as expected
(Fig. 13.10a). The presence of ciprofloxacin facilitates the adsorption of Cu(II) by
forming GH-CIP-Cu(II) ternary complex. The water confined on the adsorption sur-
face of GH can boost by the formation of adsorption sites through hydrogen bond-
ing or weaken adsorption by shielding [21].
The different oxygen contents on the adsorbent also affect the adsorption capac-
ity of ciprofloxacin. MWCNTs with different oxygen contents from 2.0 to 5.9%
show a gradual increase in normalized maximum adsorption capacity with an
increase in oxygen content (Fig.  13.10b). In the case of MWCNTs, the role of
hydrophobicity, dispersibility, and the inhibition of water cluster plays a major role
in the adsorption of ciprofloxacin. The adsorption data fit well with Dubinin–
Radushkevich as well as Langmuir adsorption isotherm model [23]. Elessawy et al.
used magnetic fullerene nanocomposite prepared from catalytic thermal dissocia-
tion of PET plastic wastes using ferrocene as a catalyst and precursor for Fe3O4. It
was a solvent-free, simple, and environmentally friendly method. The prepared
nanocomposite demonstrated a high surface area due to the presence of micropores
and mesopores, along with super paramagnetic property [25]. The strong EDA
interaction between ciprofloxacin and the fullerene nanocomposite was due to the
electron-donating effect of the amine group and F atom of ciprofloxacin with π–
electron depleted regions on the surface of the adsorbent. As shown in Fig. 13.10c,
there would be hydrogen bonds between the hydroxyl group on the adsorbent and
nitrogen-containing groups on ciprofloxacin.
Liu et  al. investigated the adsorption behavior of different sulfonamides onto
hydroxylated MWCNTs using density functional theory calculations. Molecular
structure of sulfonamides plays a crucial role in the adsorption of sulfonamides and
adsorption followed pseudo-second-order kinetics and Langmuir isotherm [26].
The fitness of the adsorption model depends on the physical and chemical charac-
teristics of adsorbent and antibiotics, apart from concentration, pH, and tempera-
ture. The adsorption data follows pseudo-second-order kinetics and Langmuir
adsorption isotherm. The frontal molecular orbital energy and dipole moment are
used as adsorption indicators for the adsorption affinity of sulfonamides onto CNTs.
13  Carbonaceous Nanomaterials for Environmental Remediation 335

Fig. 13.10  Adsorption mechanism of ciprofloxacin on wetted surface of (a) graphene hydrogel
[21], (b) MWCNT [ 23], (c) functionalized magnetic fullerene nanocomposite [25]

Fig. 13.11  Mechanism for the adsorption of sulfonamides on hydroxylated MWCNTs [26]

Figure  13.11 shows the mechanism behind the adsorption of sulfonamides on


hydroxylated MWCNTs.
Wang et  al. investigated the interaction mechanism of sulfamethoxazole with
different carbonaceous nanomaterials such as rGO, GO, graphene nanoplatelet,
MWCNT, and N-doped MWCNT to understand its adsorption behavior [27]. The
influence of dissolved humic acid (HA), aromatic compounds, and oxygen-­
containing functional groups were also tested by 1H NMR and AFM studies
(Fig.  13.12). The presence of humic acid modified the morphology of rGO and
MWCNT from linear to dispersed form based on its concentration level. It was
336 N. Sasirekha and Y.-W. Chen

observed that rGO has the strongest affinity towards sulfamethoxazole. The oxygen-­
containing functional groups on the surface of carbonaceous nanomaterials reduce
the π–π interactions with sulfamethoxazole.
Tetracycline is an organic compound with four aromatic rings containing various
functional groups. It is known to retain cationic form at all pH range. Graphene and
GO adsorb aromatic compounds through π–π stacking. The adsorption of tetracy-
cline on GO is dominated by π–π interactions due to the presence of four aromatic
rings. Irrespective of the net charge of tetracycline, the positively charged species
present orient in such a way that the positively charged group interacts with the
surface of GO through cation–π bonding, which is also responsible for adsorption.
In the case of other adsorbents like CNT and graphene also, π–π electron donor–
acceptor interactions and cation–π bonding influence the adsorption of tetracy-
cline [28].
The co-adsorption properties of arsenic (V) and tetracycline on yttrium-­
immobilized-­GO-alginate hydrogel was explored and found to show optimum per-
formance at pH  5. The 3D hydrogel showed excellent maximum adsorption
capacities of 273.39 mg g−1 and 477.9 mg g−1 for As(V) and tetracycline, respec-
tively [29]. The adsorption of As(V) was attributed to the ion exchange formed
between hydroxyl groups of the hydrogel and H2AsO42− and hydrogen bonding with
oxygen-containing groups on the hydrogel (Fig. 13.13a). However, the adsorption
of tetracycline was ascribed to electrostatic interactions, hydrogen bonding, π–π
EDA interactions, n–π EDA interactions, and cation-bonding bridge effects. The
adsorption isotherm of As(V) and tetracycline obeyed Langmuir isotherm and

Fig. 13.12  Schematic representation of the interaction between sulfamethoxazole (SMX) and gra-
phene and MWCNT and the effect of humic acid in water [27]
13  Carbonaceous Nanomaterials for Environmental Remediation 337

Freundlich isotherm, respectively. The presence of tetracycline negatively influ-


enced the adsorption of As(V) through competition for the adsorption sites. The
presence of As(V) showed contradictory effects by enhancing the adsorption of tet-
racycline through the anion–π interactions and suppressing through competition for
adsorption sites.
Li et al. used GO/TiO2 composite for adsorption of chlorotetracycline from aque-
ous solution and observed maximum adsorption capacity of 261.10 mg g−1, deter-
mined by Langmuir model [30]. The kinetics follows pseudo-second-order kinetic
model. The adsorption of chlorotetracycline may be through ligand exchange
between chlorotetracycline and TiO2, and π–π EDA interaction, hydrogen bond, and
cation–π bonding between chlorotetracycline and GO (Fig. 13.13b).
Ma and his team [22] have given a detailed review of the adsorptive removal of
antibiotic groups such as tetracyclines, sulfonamides, macrolides, quinolones from
aqueous solution using activated carbon, CNT, and graphene as adsorbents. CNT
bundles have four different adsorption sites for antibiotics such as interior site, exte-
rior site, groove site, and interstitial site. The merits of applying CNTs are the pres-
ence of abundant π–π bond between carbon atoms along with a high surface area,
thermal and chemical stability. However, the poor solubility of CNTs in water and
organic solvents limits the application. It can be overcome by employing surface
modification using covalent and non-covalent links and by filling the interior cavity.
According to literature reports, SWCNTs showed better adsorption capacities for
oxytetracycline, ciprofloxacin [31], dimetridazole, metronidazole [32], lincomycin,
and sulfamethoxazole [33] than DWCNTs and MWCNTs due to a high specific
surface area, pore volume, and basicity. The physico-chemical properties, espe-
cially surface properties, determine the absorption capacity of carbonaceous nano-
materials. The surface functionalization indeed changes the adsorption capacity due
to a decrease in the surface area after nitrogen doping and functionalization with
carboxylic acid. The presence of π–π stacking interactions enhance the adsorption
of antibiotics, which in turn is also influenced by the presence of the basic surface

Fig. 13.13  Schematic representation of the mechanism of (a) adsorption and co-adsorption of
As(V) and tetracycline by Y-GO-SA [29] (b) removal of chlorotetracycline from aqueous solution
using GO/TiO2 composite [30]
338 N. Sasirekha and Y.-W. Chen

[32]. Apart from π–π stacking interactions, electrostatic and hydrophobic interac-
tions did have a slight influence on adsorption capacity based on pH. Above pH 11,
electrostatic interactions have no influence, whereas at pH 2, electrostatic repulsion
hinders the adsorption of nitroimidazole antibiotics on CNTs. Above pH 7, hydro-
phobic interactions come into play due to a decrease in solubility. Carrales-Alvarado
et al. applied Sheindorf–Rebuhn–Sheintuch adsorption model to interpret competi-
tive adsorption of dimetridazole, metronidazole for the same adsorption sites on
SWCNT. Also, the adsorption equilibrium data for SWCNT, MWCNT, nitrogen-­
doped CNT, and MWCNT-COOH were interpreted with the Redlich–Peterson
(R–P) isotherm model.
Researchers have attempted to improve the adsorption capacity of nanocarbon
materials by modifying the surface properties and designing novel composites to
utilize the synergistic effect of different combinations. Sun et al. attempted to pre-
pare a composite consisting of Zr-based metal organic framework (MOF) and GO
for the absorption of tetracycline hydrochloride (TC) [34] and showed an absorption
efficiency of 94.88%. It was reported that the interaction between TC and GO is
primarily through π–π interaction and hydrogen bonding. However, the chemisorp-
tion of TC on UiO-66-(OH)2/GO (MOF) is based on acid–base interaction between
the amine group in TC and Zr–O cluster with partial Lewis acid properties. The
adsorption of tetracycline on the carbonaceous nanomaterials depends on the incu-
bation time, temperature, pH, and ionic strength.

Degradation of Pesticides

Pesticides play a crucial role in the improvement of crop production to meet the
growing food demand by controlling pest invasion. However, around 99.9% of the
pesticides move away from the target organisms and enter into the environment as
residue. The pesticides can be classified into herbicides, insecticides, and fungicides
based on the target organisms, of which 46% of world-wide usage accounts for
herbicide application in order to control weeds. Based on the chemical nature, pes-
ticides are classified into (i) benzoic acid, dipyridyl, phenoxy and triazines based
herbicides, (ii) carbamates, organochlorines, organophosphates and pyrethroids
based insecticides and (iii) phenylamines and phthalimides based fungicides. The
removal of pesticides through the chemical sorption method is better than biological
removal methods, although regeneration is the major constrain. The adsorption of
pesticides is through physical entrapment or chemical binding (van der Waals
forces, dipole–dipole, ion–dipole, cation exchange, strong covalent bonding) with
the adsorbents. Mojiri et al. discussed a detailed review of the removal of pesticides
from aquatic environments by adsorption methods [35].
Lima et  al. gave a theoretical proof to show the capability of C60 fullerene to
adsorb glyphosate, a potential organophosphorus herbicide in an aqueous medium
[36]. The DFT studies, fully atomistic molecular dynamics simulation and binding
energy calculations on the adsorption of pesticides such as carbaryl (insecticide),
13  Carbonaceous Nanomaterials for Environmental Remediation 339

catechol (insecticide), and fluridone (herbicide) on GO confirmed that π–π stacking


and van der Waals interactions accounted for the major adsorption interactions [37].
Mandeep et  al. used Pt-Cu decorated pyridine-like nitrogen-doped graphene to
adsorb glyphosate and DFT studies revealed that the presence of Pt-Cu clusters
enhanced the interactions between the adsorbent and adsorbate significantly [38]. In
another notable study [39], the adsorption capacity of graphene-MnFe2O4 nanocom-
posite supported on vegetal activated carbon was investigated for the adsorption of
glyphosate in a fixed-bed column. Pentachlorophenol is a potent fungicide, insecti-
cide, and bactericide with high toxicity. Ab initio simulations expressed the poten-
tial of fullerene (C60), SWCNT, and graphene as nanofilter platforms for the
adsorption of pentachlorophenol [40].
Among the herbicides, triazines are highly selective in nature and some of the
triazines are persistent in nature that can lead to mutations, birth defects and hor-
monal imbalances in human beings. In a recent study [41], porous carbon composite
was prepared from graphene-modified polymerization of high internal-phase emul-
sions and applied for the adsorption and desorption of triazine herbicides such as
simazine, prometryn, and prometon. The presence of π–π interactions and hydro-
phobic interactions between the triazine herbicides and porous carbon composite
resulted in maximum adsorption capacity of 33.4, 34.5, 33.8 μg g−1 for simazine,
prometryn, and prometon, respectively.
Fenuron is an emerging herbicide that is persistent in the environment and highly
water-soluble in nature. The MWCNT showed 90% adsorption of fenuron and the
adsorption data fitted well with Tempkin, Freundlich, Langmuir, and Dubinin–
Radushkevich models [42]. The kinetics data obeyed pseudo-first-order kinetics.
The rapid adsorption of fenuron on MWCNT was supported by negative free energy
values. Molecular docking studies demonstrated that there were different types of
hydrophobic interactions between fenuron and MWCNT such as π–σ, π–π stacked,
π–π T-shaped, and π–alkyl type. MWCNTs are also efficient in the removal of diu-
ron, a phenyl urea herbicide with adsorption of 90.5% and followed pseudo-first-­
order kinetics [43]. In the case of diazinon pesticide, MWCNT can remove 100%
from aqueous solutions at optimum concentration, adsorbent dose, and contact
time [44].
In another report [45], the adsorption capacity of oxidized MWCNT (OMWCNT)
was compared with its magnetic form (OMWCNT–Fe3O4) and composite form
(OMWCNT–κ-carrageenan–Fe3O4) for the adsorption of diquatdibromide herbi-
cide. It was noted that the adsorption and regeneration capacity of non-magnetic
OMWCNT was higher than both the magnetic forms. When compared to the maxi-
mum adsorption capacities, OMWCNT has 2.8 times higher than magnetic
OMWCNT–Fe3O4 and 5.4 times higher than magnetic OMWCNT–κ-carrageenan–
Fe3O4 (Fig.  13.14). However, all the OMWCNT forms have the potential for the
adsorption and regeneration of diquatdibromide herbicide. The adsorption iso-
therms fitted well with the Langmuir model, and it followed pseudo-second-order
kinetics. The adsorption of diquatdibromide on all the OMWCNT forms was facili-
tated by π–π and electrostatic interactions. The OMWCNT has the capacity to
regenerate and reuse for at least five times.
340 N. Sasirekha and Y.-W. Chen

Fig. 13.14  Schematic illustration of the possible interactions between OMWCNT adsorbents and
diquatdibromide herbicide and the corresponding adsorption isotherms at different tempera-
tures [45]

Nanosized biochar has also been used for the adsorption of pesticides. In recent
work [46], multifunctional β-cyclodextrin MOF-derived porous carbon was pre-
pared and used for the adsorption of amide herbicides (metolachlor, alachlor, aceto-
chlor, pretilachlor) as well as to act as a potassium nutrient source to improve,
fertility of the soil. The possible interactions between the porous carbon and amide
herbicides are through electrostatic interaction (repulsive), π–π interaction, dipole–
dipole hydrogen bonding, and Yoshida hydrogen bonding.

Degradation of Dyes

Dyes play a major role in textile, rubber, leather, tanning, paper, and printing indus-
tries. The dyes are often classified based on the chemical structure of the dyes into
nitro dyes, azo dyes, indigoid dyes, anthraquinone dyes, triarylmethane dyes, and
nitroso dyes. Based on the chemical nature, dyes can also be differentiated into acid
dyes, direct dyes, reactive dyes, basic dyes, and dispersed dyes. Another demarca-
tion can be based on the ionic charge of the dye particles into anionic (acid, direct,
and reactive dyes), cationic (basic dyes), and non-ionic (dispersed dyes) dyes. The
treatment of effluents containing dyes using conventional methods such as mem-
brane filtration, electrochemical osmosis, ion exchange, coagulation, etc. is quite
challenging. These dyes are carcinogenic, mutagenic, and teratogenic in nature
based on the chemical composition of the dyes. Due to their complex chemical
structures, biological treatment methods are also not promising. Adsorption is a
13  Carbonaceous Nanomaterials for Environmental Remediation 341

simple, cheap, sustainable, and effective method for the removal of dyes in indus-
trial wastewater, and the adsorbents can also be regenerated.
There are some critical reviews on the adsorption potential of CNTs and func-
tionalized CNTs for the removal of dye molecules from aqueous solution [47, 48].
The factors that control the adsorption of dyes are contact time between the adsor-
bent and adsorbate, pH, temperature, concentration of dye, the nature of adsorbent
and adsorbate, and the physico-chemical properties of adsorbent. The role of pH is
significant in the adsorption of Congo red and methyl orange (MO) by pristine
MWCNTs due to its influence on the surface characteristics of MWCNT. The inter-
actions involved between the dye and MWCNT are mainly through electrostatic
interactions. The adsorption capacity increases with an increase in the concentration
of the dye. From the reported literature results, it is observed that the kinetics data
were in compliance with pseudo-second-order kinetic model irrespective of the dye,
whereas adsorption isotherm obeyed isotherm models (Langmuir, Freundlich,
Temkin) based on the nature of the dye. The adsorption process is an exothermic
one and thermodynamically favored with negative free energy and a decrease in
entropy. The interactions between MWCNT and dyes such as Congo red, ponceau
4R, and allura red were investigated by combining adsorption isotherms and iso-
thermal titration microcalorimetry [49]. The results showed that the dye adsorption
was enthalpically driven and entropically unfavorable (Fig. 13.15).
Modification of CNTs by functionalization (oxidized, magnetized, polymerized,
magnetically polymerized) can enhance the adsorptive removal of dyes and easy
separation of adsorbents. The adsorption mechanism depends upon the nature and
structure of the adsorbent and the functional groups present on the surface. In gen-
eral, the interactions between the dye and CNTs are through electrostatic interac-
tions, hydrophobic interactions, van der Waals forces, hydrogen bonding, and π–π
stacking. In a recent study [50], as-grown SWCNTs (As-Fe-SWCNTs) prepared
from arc discharge method using Fe catalyst that has bounded Fe nanoparticles and
iron oxide nanoparticles anchored on SWCNT composites (AO-Fe-SWCNTs) were
used for the adsorption of methylene blue (MB) and MO dyes. AO-Fe-SWCNTs
reached equilibrium within 10  min with a maximum adsorption capacity of
256.69 mg g−1 and 93.58 mg g−1 for MB and MO, respectively (Fig. 13.16a, b).The
role of iron oxide nanoparticles was significant in that study.

Fig. 13.15  Adsorption interactions between red azo dyes and MWCNTs [49]
342 N. Sasirekha and Y.-W. Chen

GO has a negative zeta potential that facilitates easy adsorption of positive dyes
(MB), whereas different electrostatic interactions help in the adsorption of negative
dyes (MO) [51]. The N–H group present in MB and MO enables π–π interactions
between the dye and GO, and electrostatic interactions occur between =N+H group
of the dyes and oxygen functional group of GO. The ab initio molecular dynamics
studies to understand the selective adsorption behavior of MB and MO over GO
showed that MB strongly adsorbed on GO than MO. Figure 13.17a demonstrates
the most probable arrangement of the dyes on GO along with the adsorption energy.
The introduction of magnetic nanoparticles in GO has the ability to limit the issues
related to the separation and reuse of the adsorbent. There are different ferrite-based
magnetic nanoparticles that have been explored for the same. CoFe2O4 ferrite loaded
GO composite was used as an adsorbent for the removal of MB, rhodamine B
(RhB), and MO, where bare CoFe2O4 did not show any adsorption [52]. The selec-
tive adsorption behavior was of the following order: MB > RhB > MO. The adsorp-
tion process fitted well with Langmuir isotherm model and showed a superior
adsorption capacity of 355.9 mg g−1 and 284.9 mg g−1 for MB and RhB, respec-
tively. The kinetics studies data were in accordance with pseudo-second-­order
kinetics. Molecular dynamics studies revealed that π–π interactions, electrostatic
interactions, and high specific surface area influenced the adsorption of dyes on GO
(Fig.  13.17b). The defect sites and oxygen-containing functional groups also
affected the adsorption behavior of the composite. Among the presences of various
functional groups on GO, carboxyl enhanced the adsorption ability of GO than
hydroxyl or epoxy groups.
Sun et al. modified MOF (hydroxyl-modified Uio-66) framework with GO as a
guest to form a nanocomposite to adsorb MB [34]. It was found to be highly effec-
tive with an adsorption efficiency of 99.96% at optimum reaction parameters. The
adsorption isotherm fitted well with the Freundlich isotherm model and confirmed
the heterogeneous surface of the composite and multi-layer adsorption of MB. The
pH played a vital role in the adsorption of cationic dye, MB. At low pH, there was
competitive adsorption on Uio-66(OH)2/GO between excess H+ ions and MB. At
high pH, the hydroxyl groups deprotonated the surface groups of the composite and

Fig. 13.16 Adsorption of (a) MB and (b) MO on As-Fe-SWCNTs and AO-Fe-SWCNTs


(CMB = 20 mg L−1, CMO = 5 mg L−1, pH = 4, adsorbent dosage = 0.2 g L−1, T = 25 °C) [50]
13  Carbonaceous Nanomaterials for Environmental Remediation 343

Fig. 13.17 (a) Adsorption of MO and MB on GO [51], (b) Adsorption of MB, RhB and MO on
CoFe2O4/GO [52]

converted the positively charged Uio-66(OH)2/GO into negatively charged one and
enhanced the adsorption capacity of the composite. However, excess hydroxyl
groups may compete with the composite for MB and decreased the adsorption
capacity. In order to achieve maximum adsorption, a mildly alkaline medium was a
suitable one. Functionalized magnetic fullerene nanocomposites prepared from
polyethylene terephthalate bottle wastes have also been used for the adsorption of
MB and acid blue 25 (anionic) dyes through electrostatic interaction, π–π stacking
interaction and hydrogen bonding [28]. Hybrid carbonaceous nanomaterials are bet-
ter than the unmodified counterparts.

Degradation of Other Organic Pollutants

Apart from the above-discussed organic pollutants such as antibiotics, pesticides,


and dyes, polycyclic aromatic hydrocarbons (PAHs), polybrominated diphenyl
ethers (PBDEs), perfluorooctanesulfonate (PFOS), polychlorinated biphenyls
(PCBs), polychlorinated dibenzo-p-dioxins, etc. are some of the other organic
344 N. Sasirekha and Y.-W. Chen

pollutants that are present in water samples. Most of the organic pollutants are
known to be carcinogenic to humans. Based on the structure and solubility of the
organic compounds, their availability in surface water varies. PAHs are insoluble in
water and persistent in nature. Fluoranthene, phenanthrene, anthracene, and pyrene
are some of the common PAHs found in drinking water. Adsorption is a preferred
method to remove organic pollutants from wastewater or surface water. As dis-
cussed previously, carbonaceous nanomaterials have been explored for the removal
of organic pollutants and found to be highly effective.
CNTs are found to be suitable for the adsorption of organic pollutants because of
their high specific surface area, hollow and layered structures, and small size; in
particular, they exhibit strong interaction with the pollutants through van der Waals
forces, electrostatic interactions, hydrogen bonding, ion exchange, π–π EDA inter-
actions, electrophobic interactions, and micropore filling. Organic pollutants are
always bulky in nature, and hence, the preferred adsorption sites in CNTs may be
groove edges of the nanotube bundles and external surface. CNTs display signifi-
cant adsorption affinity for organic pollutants based on their chemical structure. The
selectivity and specificity can be fine-tuned by changing the surface func-
tional groups.
The adsorption property can also be evaluated by applying quantitative struc-
ture–property relationship (QSPR) modeling approach, where the requisite features
to enhance the adsorption of organic pollutants can be understood. Ghosh et  al.
applied partial least squares regression based QSPR models to understand the
physico-chemical properties and mechanism behind the adsorption of 40 synthetic
organic pollutants on SWCNTs [53]. The outcome of the work revealed that through
hydrophobic interaction, electrostatic interaction, and hydrogen bonding, the
organic pollutants may get adsorbed on SWCNTs. The organic compounds with
high hydrophobicity may adsorb easily on SWCNTs than less hydrophobic pollut-
ants, which is directly related to the size of the molecule. Figure 13.18 demonstrates
the probable adsorption mechanisms and interactions behind the adsorption of
potential organic pollutants such as naphthalene, phenanthrene, 4,6-diaminopyrimi-
dine, 2,4-dinitrotoluene, etc. on SWCNT.  In the case of phenanthrene, which is
non-ionic in nature, it has an intensive electron donor property that facilitates the
easy donation of π-electrons and gets transformed into cationic form which, in turn,
results in better adsorption on functionalized SWCNTs. Roy et al. also carried out
similar work and concluded that the adsorption of organic pollutants on SWCNTs
was based on the chemical structure of the pollutant, especially positive influence
directly related to increase in the number of aromatic rings, unsaturation, polar
groups substituted in aromatic rings, oxygen and nitrogen atoms, size, and hydro-
phobicity [54]. The negative influence on the adsorption may be due to the presence
of chlorine atoms, C–O groups, and aliphatic primary alcohols. The dispersibility of
SWCNTs can be improved by using organic polar solvents with low donor numbers,
low ionization potential, and a high degree of branching.
Bisphenol A is another persistent organic pollutant, also being considered as an
important endocrine-disrupting chemical. Ma et  al. functionalized graphene by
KOH etching and applied for the adsorption of hydrophobic organic contaminants
13  Carbonaceous Nanomaterials for Environmental Remediation 345

Fig. 13.18  Adsorption mechanism derived from QSPR model for the adsorption of synthetic
organic chemicals onto SWCNTs/functionalized SWCNTs [53]

such as naphthalene, phenol, nitrobenzene, and bisphenol A [55]. The adsorption


capacity of the functionalized graphene was found to increase by 2–8 times than the
pristine graphene dominated by hydrophobic and π–π interactions along with π–π
EDA interactions and hydrogen bonding. Mesoporous carbon material such as
CMK-3 synthesized by hard template method using SBA-15 can be used as an
adsorbent. The CMK-3 is hydrophobic in nature and has a well-defined intercon-
nected nanostructure with a uniform pore size of 3–7 nm. The large surface area and
interlinked nanorods facilitate fast intraparticle diffusion and allow more adsorption
of organic pollutants. Jeong et  al. modified CMK-3 with ethylenediamine
(MCMK-3) to selectively adsorb aromatic compounds with cyclic benzene rings
[56]. They attempted to adsorb bisphenol A onto MCMK-3 and achieved a maxi-
mum adsorption capacity of 238.01  mg  g−1 at 298  K and pH  6.4, which may be
attributed to a high surface area and hydrophobicity of MCMK-3. The interactions
between bisphenol A and MCMK-3 were due to the influence of π–π bonding,
hydrophobic and electrostatic interactions (Fig. 13.19).
346 N. Sasirekha and Y.-W. Chen

Fig. 13.19  Modified mesoporous carbon (CMK-3) for improved adsorption of bisphenol A [56]

Removal of Inorganic Metal/Metalloid Cations

The presence of inorganic metals in wastewater and drinking water has to be


removed because of their high toxicity even at low concentrations due to bioaccu-
mulation. Inorganic metals and metalloids cannot be degraded biologically and
need chemical treatment to remove from the wastewater. Chemical treatments such
as adsorption, filtration, precipitation, reduction, ion-exchange, coagulation and
flocculation, flotation, photocatalysis, and electrochemical methods are used for the
removal of heavy metals from effluents. Among the different approaches available,
adsorption is a simple and cost-effective method, without compromising efficiency
and selectivity. It can also be used at a wide pH range and suitable for large-scale
applications. The commonly available inorganic metals and metalloids in wastewa-
ter are mercury, arsenic, lead, cadmium, cobalt, chromium, zinc, copper, etc.
The adsorption interaction between carbonaceous nanomaterials and metals are
different from that of the adsorption of organic compounds. The dominant interac-
tions are ion exchange process, surface complexation, π–π interaction, electrostatic
interaction, and precipitation based on the nature of the adsorbent and adsorbate.
Besides the nature of adsorbent, the adsorption efficiency is also influenced by the
amount of adsorbent, concentration of metal ions, ionic strength, agitation speed,
contact time, temperature, and pH.
CNTs have been used for the removal of heavy metals because of their high
adsorption rate and efficiency. Fiyadh et al. have given a comprehensive review on
the role of CNTs in removing heavy metals and concluded that functionalized CNTs
with deep eutectic solvents as adsorbents could be an efficient adsorption method
[57]. Ouni et al. highlighted some of the recent developments on the adsorption of
13  Carbonaceous Nanomaterials for Environmental Remediation 347

heavy metals by CNTs, acid-modified CNTs, and functionalized CNTs. The adsorp-
tion capability of acid-modified CNTs is better than raw CNTs due to the enhanced
number of adsorption sites by the treatment. However, functionalized CNTs have
the best adsorption capability due to dispersed adsorption sites [58]. The presence
of functional groups on CNTs is crucial in deciding their adsorption behavior and
adsorption mechanisms to selectively adsorb the heavy metals [59]. The introduc-
tion of heteroatom functional groups can be achieved by oxidation, nitrogenation,
and sulfurization. According to Singha Deb et al., the introduction of sulfur ligands
on MWCNTs using sulfur, thiol, and dithiocarbamate was found to enhance the
adsorption capacity of MWCNTs toward mercury removal [60]. A recent review
discussed the role of CNTs, graphene, GO, and rGO for the adsorptive removal of
Pb(II) ions [61]. The interaction between Pb(II) ions and adsorbent were mainly
electrostatic interactions and the influence of pH and ionic strength played a pre-
dominant role. The adsorption isotherm fitted well with Langmuir adsorption model
and the kinetics data with pseudo-second-order, which confirmed monolayer
adsorption of Pb(II) ions. In the case of CNTs, when the pH is higher than the iso-
electric point of CNT, the adsorption of metal ions onto CNT increases because of
increased electrostatic interaction between the negatively charged CNTs and metal
cations [62]. These electrostatic interactions can also be increased by introducing
heteroatoms with lone pair of electrons. The desorption of metals takes place at
lower pH values. The surface complexation with metal ions is facilitated at the che-
lating sites present at the edges of CNTs. Apart from these factors, the physico-­
chemical properties of CNTs, defects, cavities, and nanopores also influence the
adsorption capacity of CNTs. CNT modified with poly-amidoamine dendrimer (G4)
was used to adsorb Cu2+ and Pb2+ ions from industrial wastewater and showed a
maximum adsorption capacity of 3333 mg g−1and 4870 mg g−1, respectively [63].
MWCNTs decorated with alumina exhibited an adsorption capacity of 27.21 mg g−1
toward the adsorption of Cd(II) ions through electrostatic interactions, hydrogen
bonding, and protonation of Al2O3 as shown in Fig. 13.20a [64].

Fig. 13.20  Adsorption interactions between (a) Cd(II) ions with Al2O3/MWCNTs [64] and (b)
Cr(VI) ions with polyethyleneimine cross-linked GO [65
348 N. Sasirekha and Y.-W. Chen

DFT studies on the adsorption mechanism of graphene and GO for the removal
of cadmium and lead revealed that the heavy metals adsorbed strongly on GO than
pristine graphene due to the presence of oxygen functional groups [65]. The cad-
mium adsorption was through dispersive forces; however, lead adsorption took
place through dispersive forces as well as by covalent bonding, as lead can act as
electron donor. Grafting of GO with dialdehyde cellulose using triethylenetetramine
as cross-linking reagent showed excellent performance in removing Cu2+ and Pb2+
ions at pH 5 with a maximum adsorption capacity of 65.1 mg g−1and 80.9 mg g−1,
respectively [67]. GO terminated hyperbranched amino polymer-carboxymethyl
cellulose ternary nanocomposite was used for the adsorption of Cu2+ and Pb2+ ions
in a fixed-bed column and showed a maximum adsorption capacity of 137.48 mg g−1
and 152.86 mg g−1, respectively [68]. The high performance was attributed to the
presence of porous structure along with a large number of unsaturated atoms and
heteroatom functional groups at the surface. The DFT studies revealed the prefer-
ence of binding of each functional group toward the heavy metals. The functional
groups such as –OH, –NH2, and –COOH strongly preferred to bind with Pb2+,
whereas –CONH– did not show a specific preference toward Pb2+ or Cu2+. The per-
formance was also governed by complexation and ion exchange of metal ions with
N-containing functional groups and O-containing functional groups, respectively.
Polyethyleneimine cross-linked GO exhibited an adsorption capacity of
436.20  mg  g−1 for the removal of Cr(VI) ions [65]. The dominant interactions
between them were electrostatic interaction, chelation, and reduction processes
(Fig. 13.20b). Ahmad et al. have explained the role of functional groups present in
GO for the adsorption of cationic and oxyanionic heavy metals along with their
interactions [69].
The application of fullerenes and graphene quantum dots in the removal of heavy
metals has not been explored extensively. Abdelsalam et al. used DFT studies to
understand the adsorption capacity of graphene quantum dots to remove hydrated
heavy metals such as Cd and Pb [70].

13.3.2  Photocatalysis

Heterogeneous photocatalysis is one of the successful approaches used for the deg-
radation of organic pollutants and metal reduction from wastewater. In a photocata-
lytic process, a semiconductor when exposed to light with energy equal to or greater
than its bandgap absorbs the energy of the photons and the electrons get excited
from the valence band to the conduction band, generating electron–hole pairs. The
excited electrons and holes migrate to the surface of the semiconductor and initiate
the oxidation–reduction reactions to degrade pollutants. Some of the common semi-
conductors as photocatalysts are TiO2, ZnO, Fe2O3, CdS, CdSe, etc. The main chal-
lenge in executing a photocatalytic reaction is to minimize electron–hole
recombination as well as to fabricate an efficient photocatalyst that absorbs light in
the visible region.
13  Carbonaceous Nanomaterials for Environmental Remediation 349

Photocatalysis is a surface phenomenon, where the surface area plays a major


role in deciding the adsorption of pollutants on the active sites of the photocatalyst.
Nanomaterials are quite promising in photocatalytic reactions because of their high
surface area and in turn more active sites for adsorption. The detailed principle,
mechanisms, and applications of photocatalysts are discussed in another chapter.
The photocatalytic efficiency and adsorption performance of a photocatalyst can be
improved by fabricating a hybrid or composite photocatalyst in order to exploit the
synergic effects of the materials or catalysts combined.
CNTs, graphene, fullerene, and carbon quantum dots can be used as co-catalysts
for photocatalytic degradation of pharmaceutical organic pollutants, dyes, pesti-
cides, and other organic pollutants. They are found to have a positive influence on
the photocatalytic efficiency of photocatalysts. The high electron transport and
reception ability of CNTs make it an attractive co-catalyst for photocatalytic degra-
dation reactions. AgCl/CNTs/graphitic carbon nitride (g-C3N4) nanocomposite was
investigated for the degradation of tetracycline under visible light and found to dis-
play a photocatalytic efficiency of 86.44% [71]. The composite could also be used
as an antibacterial agent (Fig.  13.21a). CNTs modified g-C3N4/BiVO4 composite
was used as a photocatalyst for the photocatalytic degradation of phenol and found
to be capable of removing 80.6% [68]. The kinetics data followed pseudo-first-order
kinetics and obeyed the Temkin adsorption model.
Apart from being a co-catalyst, graphene can also be used as a support for pro-
viding high dispersion of the semiconductors. It can also provide an effective chan-
nel for electron transport. Fe2O3/graphene/CuO hybrid material with a bandgap of
1.82 eV showed remarkable photocatalytic efficiency for the degradation of methy-
lene blue, even in the absence of other oxidants (Fig. 13.22a). The magnetic prop-
erty of the hybrid material assisted an easy separation of the photocatalysts by
applying an external magnetic field [73]. Pan et al. have given a detailed review of
the properties, mechanism, synthesis, and applications of fullerene C60-based pho-
tocatalysts [75]. The presence of fullerene in composite influences the electron
transfer process in the semiconductor composite by acting as an electron donor or
acceptor, electron transfer mediator, and electron donor and acceptor. The incorpo-
ration of C60 into g-C3N4 matrix improved the photooxidation capability of g-C3N4
by lowering the valence band energy position [74]. C60/g-C3N4 displayed high

Fig. 13.21  Schematic representation for the photocatalytic degradation of (a) antibiotics by AgCl/
CNTs/g-C3N4 nanocomposite [71] (b) phenol by g-C3N4/CNT/BiVO4 nanocomposite [72]
350 N. Sasirekha and Y.-W. Chen

Fig. 13.22  Schematic representation of mechanism of charge separation and photocatalytic deg-
radation of (a) methylene blue by Fe2O3/graphene/CuO nanocomposites [73], (b) organic pollut-
ants by C60/g-C3N4 [74]

photocatalytic performance toward the degradation of phenol and methylene blue


due to the formation of conjugative π bond between fullerene and carbon nitride
(Fig. 13.22b).

13.3.3  Membrane

Carbonaceous nanomaterials can also be fabricated into membranes for the applica-
tion of desalination and wastewater treatment. Desalination is an important process
for the production of freshwater from seawater, wastewater, and saline water to meet
the growing potable water demand. In this process, the concentration of dissolved
salts and solids will be removed by thermal processes or desalination membranes.
Membrane-based technologies outperform thermal processes due to low energy
consumption. In general, pressure-driven processes such as microfiltration, ultrafil-
tration, nanofiltration, and reverse osmosis differ with respect to their pore size and
rejection mechanism. Membranes used for microfiltration, ultrafiltration, and nano-
filtration have pore sizes ranging from 0.1 to 10  μm, 2.0–100  nm, and <1.0  nm,
respectively. They use sieving and the basic size exclusion principle for retaining
non-dissolvable matter, macromolecules, small organic molecules, and multivalent
ions based on the pore size of the membrane. Reverse osmosis membranes are non-­
porous in nature and can remove even monovalent ions at high pressures. They work
based on diffusion mechanisms by overcoming osmotic pressure. The choice of
solute depends upon its solubility and diffusivity in the membrane. The membrane
should be able to withstand chemical and microbial attack, and high-salt-rejection
rate, high permeability, and water flux. The conventional cellulose acetate-based
membranes lack high water flux and resistance toward microbial attack. Even
though different membranes have been attempted and modified for more than six
decades, a suitable and sustainable membrane to meet the demands of water supply
13  Carbonaceous Nanomaterials for Environmental Remediation 351

has not been identified. The main challenge is to fabricate a membrane cost-­effective
that can withstand the membrane fouling, chemical degradation by oxidants, and
permeability–selectivity trade-off.
Membrane engineering attempts to manipulate synthesis approaches, surface
modifications, and techniques for the preparation of conventional membranes, espe-
cially for sub-nanometer separation. New materials such as zeolites, metal organic
frameworks, metal oxide nanoparticles, CNTs, and graphene have also been inves-
tigated, in order to fulfill the most energy-intensive angstrom level separation opera-
tion. Different approaches have been attempted to develop highly selective
membranes such as biomimetic aquaporin-based membranes (supported membrane
layers, vesicle encapsulated membranes), nanozeolite-based membranes, and self-­
assembled materials-based membranes. Recently low-dimensional materials (LDM)
membrane has attracted attention for desalination and gas separations because of
their extraordinary mass transport properties due to nanoconfinement in one or
more dimensions. Carbonaceous nanomaterials such as GO, graphene sheets, and
CNTs are some of the potential materials for membrane fabrication due to their high
mechanical strength and stability.
Graphene is a single-atom-thick material; hence it is a suitable candidate that can
be made into a membrane with minimum thickness and maximum permeability.
Pristine graphene with a high specific surface area and impermeable barrier cannot
be used as a membrane. Hence, it is necessary to introduce uniform and precise
pores in order to transform pristine graphene into a nanoporous graphene mem-
brane. The challenge lies in introducing angstrom level precise and uniform pores
on pristine graphene. Different approaches such as (1) the use of naturally occurring
pores in graphyne, (2) use of 2D slits instead of 1D pores, (3) layered or overlapping
graphene membrane have been attempted. After introducing the pores, the carbon
atoms at the edge of the pores of the nanoporous graphene membrane need to be
functionalized by either hydrophobic (hydrogen, alkyl, aryl) or hydrophilic (–OH,
substituted or unsubstituted amino) groups. The presence of the hydroxyl group
rapidly increases the water reflux, but at the expense of salt rejection performance.
The contrary holds good for the presence of the hydrophobic group, where salt
rejection improves with restriction in water flow. The functionalization also permits
ion selectivity that makes graphene membrane quite promising than reverse osmo-
sis, where all the cations and anions will be removed [76].
In the case of multilayer graphene, apart from the influence of pore size and
chemical functionalization, the distance of interlayer spacing between the graphene
sheets also affects the molecular sieving and separation. According to the latest
report [77], salt rejection and water reflux would be affected by the distance of inter-
layer spacing, width of gap, offset, and number of gaps and layers on the perfor-
mance of multilayer graphene membrane for seawater desalination derived from
molecular dynamics simulation studies. The interlayer distance greater than 0.8 nm
decreased salt rejection value. Under the reported optimal desalting conditions, the
multilayer graphene membrane showed excellent salt rejection of Mg2+ and Ca2+
ions and, in turn, confirmed its potential for seawater desalination. Most of the
investigations on graphene membranes are by molecular dynamics simulations. So
352 N. Sasirekha and Y.-W. Chen

far, different computational methods revealed the potential of the graphene mem-
brane for desalination. But experimental constraints in fabricating graphene mem-
brane limit its industrial applications.
According to literature reports, the fabrication of the GO membrane is easier
than graphene and CNT membranes. On the other hand, GO membrane shows poor
separation performance, but graphene and CNT membranes display high permea-
bility as the permeability of a membrane is inversely related to membrane thickness
[78]. The pore size determines the variation in water flux. At a larger diameter,
graphene transports higher water flux than CNT.  On the contrary, CNT delivers
higher water flux than graphene at smaller pore sizes. Similar to graphene, pristine
CNT cannot be used as a membrane as it tends to aggregate and shows lower water
flux. CNTs can be functionalized at the tip or at the core. CNTs can also be used to
encapsulate fullerene, and it was studied by molecular dynamics simulation [79].
The results were quite promising to apply as water nanochannel by displaying 100%
salt rejection.
In order to overcome the shortcomings of reverse osmosis, progressive research
work is also going on to replace reverse osmosis by forward osmosis, capacitive
deionization, membrane distillation, ultrafiltration, and nanofiltration techniques,
but it is still in the early level of investigation with a lot of challenges.
Membrane distillation is a desalination technology combining both membrane
separation and distillation process, where the difference in vapor pressure on both
sides of the membrane facilitates the separation of volatile components. Based on
the vapor collection method, it is classified into direct contact membrane distilla-
tion, air gap membrane distillation, sweep gas membrane distillation, and vacuum
membrane distillation. In general, hydrophobic polymers such as polytetrafluoro-
ethylene, polypropylene, and polyvinylidenefluoride are used as membranes. Recent
reports show the potential of some of the new materials like ceramics, CNTs, gra-
phene, and GO for the membrane distillation process. In comparison with graphene,
GO exhibits interlayer nano-channel that limits the passage of large contaminants
and has relatively simple synthesis procedures. Furthermore, fast radial heat transfer
and facile hydrophobic modification makes it a suitable material for this process.
Mao et al. attempted to modify GO laminar membrane using SiO2 nanoparticles in
order to improve surface-roughness and laminating spacing. The hydrophobicity of
the material was achieved by grafting with long alkyl chain. The modified novel
hydrophobic GO-based membrane intercalated with SiO2 nanoparticles were used
for water desalination using vacuum membrane distillation method by varying treat-
ment temperature, feed conditions, and organic humic acid. The membrane demon-
strated excellent salt rejection of 99.99% (NaCl), stabilized water flux of
13.59 kg m−2 h−1, and good antifouling ability toward surfactant and organics [80].
Hung et al. developed a method for the preparation of amphiphilic GO by stacking
flexible GO layers on modified polyacrylonitrile substrates using pressure-assisted
self-assembly method. The composite membrane showed excellent performance for
the separation of isopropyl alcohol/water mixture. When the water molecule inter-
acts with the hydrophilic hydroxide edges, it adsorbs first and diffuses rapidly
through the hydrophobic carbon core and creates a water passage channel, thereby
13  Carbonaceous Nanomaterials for Environmental Remediation 353

Fig. 13.23  Modified CNT membranes for different desalination techniques [82]

promoting high permeation flux. The movement of water molecules through the GO
creates a monolayer structure of the accumulated water molecules, and pushes the
layers away from each other and thus enlarges the d-spacing between the layers [81].
Due to its outstanding mechanical properties, CNTs are used in reverse osmosis,
forward osmosis, membrane distillation, capacitive deionization, nanofiltration, and
ultrafiltration techniques using a suitable surface modification to meet out the speci-
fications of each technique (Fig. 13.23).

13.4  C
 arbonaceous Nanomaterials for Removal
of Gas Pollutants

Air pollution accounts for nearly 4.2 million deaths worldwide due to outdoor air
pollution [82]. Air pollutants are volatile organic compounds (VOCs), inorganic
gases (NOx, SOx, carbon oxides, ozone, dioxins, PCBs, etc.), and suspended particu-
late matters. VOC can be removed (1) by modifying the process and equipment; (2)
by destruction techniques such as thermal oxidation, catalytic oxidation, reverse
flow reactor (RFR), and biofiltration; and (3) by recovery techniques using absorp-
tion, adsorption, condensation, and membrane separation.
Nanomembranes are used to filter different pollutants from exhaust fumes.
Cryogenic distillation, absorption, and adsorption are some of the techniques that
are available for gas separations. However, membranes are considered to be com-
pact and energy-efficient compared to the above techniques. The main challenge
with the conventional polymeric membranes lies in satisfying the desired high per-
meability and selectivity. Different materials have been investigated for fabricating
a membrane with high permeability and high rejection. LDM membranes such as
GO, graphene, and CNT membranes show promising results in gas separations from
the refinery, petrochemical, and natural gas mixtures. It is viable to separate hydro-
gen, oxygen, and carbon dioxide from the air as well as gas mixtures.
Adsorption is a cost-effective process to remove air pollutants and particulate
matter from the exhaust gas. Graphene-based nanomaterials are used as adsorbents
for the removal of VOCs. In general, aromatic VOCs such as benzene, toluene, and
xylene tend to adsorb on the hydrophobic surfaces of graphene. Based on the hydro-
phobic nature, the adsorption of VOCs on graphene-based materials is in the order
of rGO > pristine graphene > GO [84]. The rGO has more preference to form π–π
354 N. Sasirekha and Y.-W. Chen

bonds with the aromatic VOCs. The adsorption capacity of these materials can be
improved by increasing the specific surface area using activation. Molecular simula-
tion studies revealed the suitability of modified CNTs for the adsorption of gas
molecules. Fe-doped vacancy-defected CNTs were used for the adsorption of CO
[85] and SO2 [86] molecules, whereas functionalized CNTs (hydroxyl, carboxyl,
cyclodextrin) were found to be suitable for the adsorption of H2S molecules. In the
case of H2S adsorption on cyclodextrin-functionalized CNTs, dispersion and elec-
trostatic interactions were dominant [87].
Photocatalytic degradation is an effective method for the removal of gaseous pol-
lutants even at trace levels. Carbonaceous nanomaterials can be used as a support
for the dispersion of active semiconductors. They can provide adsorption sites for
the gaseous pollutants to adsorb for subsequent degradation and can also modify the
photogenerated charge-transfer process. Most of the degradation kinetics followed
Langmuir–Hinshelwood equation. Based on the physico-chemical properties of the
carbon nanomaterials, the photocatalytic efficiency varies. In the case of CNTs, the
helicity and the diameter of the CNTs decide their conducting property (metallic,
semiconductor). On the other hand, graphene is a conductor with zero bandgap that
needs special treatment to convert the hydrophobic graphene into hydrophilic to
deposit metal oxides on its surface to execute adsorption of gaseous molecules and
subsequent photocatalytic degradation. Modification of graphene by doping and
functionalization improves the adsorption-reactive sites on the surface. The degra-
dation efficiency was quite prominent for rGO due to the strong interactions with
the photocatalyst and less mass transfer limitations. In order to achieve high photo-
catalytic performance, the physico-chemical properties of the carbon nanomaterials
such as surface area, porosity, active sites, hydrophobicity, hydrophilicity, pore vol-
ume, light absorption, charge-transfer, etc. should also be considered. The reaction
parameters such as temperature, light irradiation, moisture, and concentration of
pollutants have to be optimized. The VOCs tend to adsorb through van der Waals
interactions, hydrogen bonding, hydrophobic effect, and π–π bonding [88]. Zhang
et al. attempted to fabricate a novel monolithic protonated g-C3N4/GO aerogel for
the photocatalytic removal of nitric oxide to nitrate with a removal ratio of 46.1%.
The positive protonated g-C3N4 combined with the negative GO and enhanced the
photocatalytic activity of the aerogel to oxidize even the secondary pollutant nitro-
gen dioxide to nitrate [89].
Both the adsorption and photocatalytic degradation of gaseous pollutants have
their own limitations. In order to enhance the degradation efficiency, both adsorp-
tion and photocatalytic degradation can be integrated. The integrated adsorptive and
photocatalytic method uses the adsorption process to adsorb the pollutants from the
gas phase to the solid phase effectively followed by photocatalytic degradation of
the adsorbed pollutants to complete mineralization.
13  Carbonaceous Nanomaterials for Environmental Remediation 355

13.5  Carbonaceous Nanomaterials for Soil Remediation

One of the important soil contaminants that need immediate attention is metalloids
such as arsenic, mercury, chromium, lead, zinc, etc. The main sources are mining
activities, industrial processes, application of contaminated domestic sludge, coal
burning, and agrochemicals. The toxic level of metalloids depends upon its chemi-
cal form. For example, arsenite [As(III)] is more toxic than arsenate [As(V)], and
organic mercury is more toxic than other forms of mercury. In soil, metalloids
because of their mobile nature may be taken up by plants or transport to groundwa-
ter along with soil water. Some of the ways to immobilize metalloids are through
sorption, precipitation, and complexation methods that depend on the soil properties
and environmental factors.
Organic contaminants such as pesticides, polycyclic aromatic hydrocarbons,
polychlorinated biphenyls, organic solvents, and pharmaceuticals enter into the soil
through anthropogenic activities. Most of them are toxic and persistent in soil and
can undergo biomagnification to affect higher tropic levels. Carbon nanomaterials
such as CNTs, graphene, and fullerene are some of the promising materials for soil
remediation because of their high adsorption capacities, surface area, and hydro-
phobicity. Because of high hydrophobicity, CNTs prefer to adsorb hydrocarbons
than alcohol. Based on the nature of the soil, the organic pollutants have variable
affinity toward the soil.
The source of antibiotics in the soil is mainly because of the incomplete metabo-
lism of antibiotics in humans as well as animals, and their discharge to the waste
stream. Eventually, the disposal of sewage sludge, poultry and livestock wastes, and
discharge of municipal wastewater into the soil accounts for the high concentration
of antibiotics in soil. The presence of antibiotics in soil disturbs the homeostasis of
soil by inhibiting the growth of soil microorganisms, decreasing soil respiration,
changing the turnover rate of soil nutrients, biogeochemical cycles, and specific
enzyme activity. Some of the highly sensitive microorganisms may disappear, and
the development of soil antibiotic-resistant bacteria and pathogenic antibiotic-­
resistant bacteria may emerge as a result. Even the sub-lethal level of antibiotics can
increase the mutation rate and resistance genes in soil. Notable antibiotics that are
found in soil are sulfonamides, tetracyclines, amoxicillin, oxytetracycline, cipro-
floxacin, tylosin, cephalosporins, etc. High concentrations of antibiotics can affect
the growth of plants and made their way to the edible part of the crop. Hence, it is
mandatory to find a method to significantly reduce the concentration of antibiotics
in soil.
In soil, the degradation or transformation of antibiotics starts immediately upon
application to soil, based on the nature and concentration of antibiotics, biotic, and
abiotic conditions. The fate of antibiotics widely depends upon their physico-­
chemical properties and molecular structure. One of the important abiotic factors is
hydrolysis, which plays a prominent role in the degradation of β-lactams, whereas
it has an insignificant role in macrolides and sulfonamides. Another abiotic factor
that has a significant role in the degradation of antibiotics is photodegradation. The
356 N. Sasirekha and Y.-W. Chen

physico-chemical characteristics of soil and climatic factors also determine the


degree of antibiotics degradation. The sorption coefficient is an important parameter
that decides the persistence of antibiotics and degradation rate in soil. Based on the
presence of organic and inorganic soil components, the antibiotics tend to form
complexes or bind strongly with soil components. When the sorption coefficient of
antibiotics in the soil is less than 15  L/Kg, the antibiotics can easily move and
degrade fast in soil, whereas the sorption coefficient greater than 4000 L/Kg leads
to highly persistent and stable residues of antibiotics. It has been reported that due
to high affinity toward soil components, antibiotics such as tetracyclines, macro-
lides, fluoroquinolones, and sulfonamides form stable residues. The pH of the soil
also influences the adsorption and desorption characteristics of antibiotics. An
increase in pH decreases adsorption of sulfonamides, due to the transformation of
cationic form to neutral and anionic form. The main sorption mechanism in the case
of the cationic form of antibiotics is electrostatic interactions through surface com-
plexation, cation exchange, and cation bridging sorption. The half-life or DT50
indicates the persistent nature of antibiotics in different soil, which depends upon
the organic carbon content in soil. Apart from abiotic degradation, microbial degra-
dation of antibiotics by soil microbes such as Microbacterium, Ochrobactrum,
Labrys etc. are also significant.
Carbon-based materials are found to be biocompatible and act as a good sorbent
for the removal of toxic substances from the environment. Graphene oxide has the
capacity to bind with antibiotics, especially tetracycline and sulfamethoxazole. Zou
et al. investigated the role of GO in inhibiting sulfamethoxazole uptake by bacteria
and the transfer of antibiotic resistance genes (ARG) among microorganisms. GO
readily forms a complex with sulfamethoxazole (GO-SMZ) and hinders the uptake
of antibiotics in bacteria. The non-covalent combination of GO-ARG alters the
properties of ARG [90].
Biochar is a solid carbonaceous material that originates from pyrolysis of bio-
mass in the presence of insignificant amount of oxygen, with a pyrolytic tempera-
ture above 250 °C. It is considered to be a cheap and environmental-friendly material
that has been used for carbon sequestration as well as for the improvement of soil
quality because of its large specific surface area, surface functional groups, and
porous structure. It is a good sorbent for the removal of pollutants from wastewater,
soil and air. However, the physico-chemical properties of biochar vary with respect
to the composition and chosen biomass.
Carbonaceous nanomaterials have proved to be a good adsorbent, photocatalyst,
and membrane filter for various environmental applications. In contrast, the fate of
carbon nanomaterials is also a major concern as it is inevitably entering the environ-
ment. Due to the high stability of these materials, their degradation in nature is quite
uncertain. The main degradation pathways are chemical degradation, photodegrada-
tion, and biodegradation. CNTs, graphene, and fullerene are reported to have cyto-
toxicity and found to influence oxidative stress response, mechanical damage, and
biological enzymes. It is noted that at low concentrations, carbon nanomaterials
have a positive influence on microorganisms. However, at high concentrations, they
affect microbes by destroying the electrostatic equilibrium or cell membrane and
13  Carbonaceous Nanomaterials for Environmental Remediation 357

organelles of microbial cells. Moreover, there will be changes in lipids and proteins
of microbial cells due to the formation of oxidative stress, which often results in
abnormal metabolism of microorganisms and further cell damage [91]. Similarly,
low concentrations do enhance the vegetative growth and crop yield, by enhancing
water uptake and transport, seed germination and growth, physiological activities,
antioxidant activities, activating water channel proteins, and promoting nutrition
absorption, whereas high concentrations reverse the effect [92]. The beneficial
effects depend on the type, physical characteristics, and concentration of carbon
nanomaterials along with the type of vegetation, application method, etc. In general,
they have shown toxicity towards animals, human beings, plants, and microorgan-
isms, as it is difficult to identify the thin line of demarcation between its beneficial
and harmful concentrations. The studies on the toxicity of carbonaceous nanomate-
rials are in vivo and in its early stages, which needs a substantial amount of work to
understand the degradation mechanism and toxicity of these materials [93].

13.6  Conclusion

Carbonaceous nanomaterials have proven their ability to remove various pollutants


from wastewater, air, and soil by means of adsorption, photocatalysis, and mem-
brane technology. The physico-chemical properties of carbon-based nanomaterials
can be engineered through functionalization or by combining different materials to
form nanocomposites or nanohybrids in order to improve their performance based
on the target pollutant. Carbonaceous nanomaterials like CNTs, graphene, GO,
rGO, fullerene, etc. are ideal for adsorbing both organic pollutants and inorganic
metals. The high specific surface area, adsorption capacity and the flexibility to tune
the physico-chemical properties make these nanomaterials a competent adsorbent.
The interactions with the adsorbent vary with the target pollutant. Organic pollut-
ants (antibiotics, pesticides, dyes) in water interact with the carbon nanomaterials
through van der Waals forces, π–π interactions, hydrophobic interactions, and elec-
trostatic interactions, whereas inorganic metals use electrostatic interactions, ion
exchange, and surface complexation, dissolution, co-precipitation, and surface pre-
cipitation to get adsorb. Adsorption isotherms are studied to understand the interac-
tion between the adsorbent and adsorbate, which helps to design an effective
treatment system based on the interpretations of the capacity of adsorbents, surface
properties, adsorption mechanism, and adsorption phenomenon. Adsorption kinet-
ics reveals the dynamics of the adsorption process by providing information about
the rate of adsorption and time taken to attain equilibrium along with probable
adsorption pathways and mechanisms. The pseudo-second-order kinetics holds
good for most of the adsorption of organic pollutants. Gaseous pollutants are
removed by adsorption, photocatalysis, and membrane filtration. Integrating both
adsorption and photocatalysis techniques provides a promising technology to
degrade gaseous pollutants efficiently. Soil reclamation by employing carbon-based
358 N. Sasirekha and Y.-W. Chen

nanomaterials needs in-depth study to understand the fate of the adsorbents in


the soil.
In order to design a novel adsorbent, it is essential to understand the interactions
between the adsorbent and adsorbate. Some of the current research focused on
investigating at a microcosmic level using computational tools in order to screen a
wide range of pollutants. From a practical standpoint, the difficulty encountered in
synthesizing these nanomaterials in large scale economically to meet out the demand
in real industries limits their applications. There is still a wide gap between the labo-
ratory scale to real commercialization and practical applications. The prime focus of
research should be on fabricating a commercially viable carbon-based hybrid mate-
rial with recycling ability to make it economically cheap and efficient.

References

1. Iijima, S., Yudasaka, M., Yamada, R., Bandow, S., Suenaga, K., Kokai, F., & Takahashi,
K. Nanoaggregates of single-walled graphitic carbon nano-horns. Chem. Phys. Lett. 1999,
309, 165–170. https://doi.org/10.1016/S0009-2614(99)00642-9
2. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. https://doi.
org/10.1038/354056a0
3. Novoselov, K. S., Geim, A. K., Morozov, S. V., Jiang, D., Zhang, Y., Dubonos, S. V., Grigorieva,
I. V., & Firsov, A. A. Science 2004, 306, 666–669. https://doi.org/10.1126/science.1102896
4. Xu, X. Y., Ray, R., Gu, Y. L., Ploehn, H. J., Gearheart, L., Raker, K., and Walter A. Scrivens.
Electrophoretic analysis and purification of fluorescent single-walled carbon nanotube frag-
ments. J. Am. Chem. Soc. 2004, 126, 12736–12737. https://doi.org/10.1021/ja040082h
5. Sun, Z., Ikemoto, K., Fukunaga, T. M., Koretsune, T., Arita, R., Sato, S., & Isobe, H. (2019).
Finite phenine nanotubes with periodic vacancy defects. Science, 363, 151–155.
6. Georgakilas, V., Perman, J. A., Tucek, J., & Zboril, R. (2015). broad family of carbon nanoal-
lotropes: Classification, chemistry, and applications of fullerenes, carbon dots, nanotubes, gra-
phene, nanodiamonds, and combined superstructures. Chemical Reviews, 115(11), 4744–4822.
https://doi.org/10.1021/cr500304f.
7. Li, Z., Liu, Z., Sun, H., & Gao, C. (2015). Superstructured assembly of nanocarbons:
Fullerenes, nanotubes, and graphene. Chemical Reviews, 115(15), 7046–7117. https://doi.
org/10.1021/acs.chemrev.5b00102.
8. Mochalin, V.  N., Shenderova, O., Ho, D., & Gogotsi, Y. (2012). The properties and appli-
cations of nanodiamonds. Nature Nanotechnology, 7(1), 11–23. https://doi.org/10.1038/
nnano.2011.209.
9. Xu, Q., Li, W., Ding, L., Yang, W., Xiao, H., & Ong, W. J. (2019). Function-driven engineer-
ing of 1D carbon nanotubes and 0D carbon dots: Mechanism, properties and applications.
Nanoscale, 11(4), 1475–1504. https://doi.org/10.1039/c8nr08738e.
10. Xu, J., Cao, Z., Zhang, Y., Yuan, Z., Lou, Z., Xu, X., & Wang, X. (2018). A review of func-
tionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation,
application, and mechanism. Chemosphere, 195, 351–364. https://doi.org/10.1016/j.
chemosphere.2017.12.061.
11. Mojica, M., Alonso, J. A., & Méndez, F. (2013). Synthesis of fullerenes. Journal of Physical
Organic Chemistry, 26(7), 526–539. https://doi.org/10.1002/poc.3121.
12. Scott, L. T. (2004). Methods for the chemical synthesis of fullerenes. Angewandte Chemie—
International Edition, 43(38), 4994–5007. https://doi.org/10.1002/anie.200400661.
13  Carbonaceous Nanomaterials for Environmental Remediation 359

13. Kim, S., Park, C. M., Jang, M., Son, A., Her, N., Yu, M., Snyder, S., Kim, D. H., & Yoon,
Y. (2018). Aqueous removal of inorganic and organic contaminants by graphene-based
nanoadsorbents: A review. Chemosphere, 212, 1104–1124. https://doi.org/10.1016/j.
chemosphere.2018.09.033.
14. Yin, Z., Cui, C., Chen, H., Duoni, Yu, X., & Qian, W. (2019). The application of carbon nano-
tube/graphene-based nanomaterials in wastewater treatment. Small, 1902301, 1–16. https://
doi.org/10.1002/smll.201902301.
15. Baig, N., Ihsanullah, Sajid, M., & Saleh, T. A. (2019, October). Graphene-based adsorbents
for the removal of toxic organic pollutants: A review. Journal of Environmental Management,
244, 370–382. https://doi.org/10.1016/j.jenvman.2019.05.047.
16. Thakur, K., & Kandasubramanian, B. (2019). Graphene and graphene oxide-based composites
for removal of organic pollutants: A review. Journal of Chemical and Engineering Data, 64(3),
833–867. https://doi.org/10.1021/acs.jced.8b01057.
17. Zhuang, Y., Yu, F., & Ma, J. (2015). Enhanced adsorption and removal of ciprofloxa-
cin on regenerable long TiO2 nanotube/graphene oxide hydrogel adsorbents. Journal of
Nanomaterials, 2015, 675862. https://doi.org/10.1155/2015/675862.
18. Khan, N. A., Najam, T., Shah, S. S. A., Hussain, E., Ali, H., Hussain, S., Shaheen, A., Ahmad, K.,
& Ashfaq, M. (2020). Development of Mn-PBA on GO sheets for adsorptive removal of cipro-
floxacin from water: Kinetics, isothermal, thermodynamic and mechanistic studies. Materials
Chemistry and Physics, 245, 122737. https://doi.org/10.1016/j.matchemphys.2020.122737.
19. Peng, X., Hu, F., Huang, J., Wang, Y., Dai, H., & Liu, Z. (2016). Preparation of a graphitic
ordered mesoporous carbon and its application in sorption of ciprofloxacin: Kinetics, isotherm,
adsorption mechanisms studies. Microporous and Mesoporous Materials, 228, 196–206.
https://doi.org/10.1016/j.micromeso.2016.03.047.
20. Veclani, D., & Melchior, A. (2020). Adsorption of ciprofloxacin on carbon nanotubes: Insights
from molecular dynamics simulations. Journal of Molecular Liquids, 298, 111977. https://doi.
org/10.1016/j.molliq.2019.111977.
21. Ma, J., Xiong, Y., Dai, X., & Yu, F. (2020a, July). Coadsorption behavior and mechanism of
ciprofloxacin and Cu(II) on graphene hydrogel wetted surface. Chemical Engineering Journal,
380, 122387. https://doi.org/10.1016/j.cej.2019.122387.
22. Yu, F., Li, Y., Han, S., & Ma, J. (2016a). Adsorptive removal of antibiotics from aqueous
solution using carbon materials. Chemosphere, 153, 365–385. https://doi.org/10.1016/j.
chemosphere.2016.03.083.
23. Yu, F., Sun, S., Han, S., Zheng, J., & Ma, J. (2016b). Adsorption removal of ciprofloxacin
by multi-walled carbon nanotubes with different oxygen contents from aqueous solutions.
Chemical Engineering Journal, 285, 588–595. https://doi.org/10.1016/j.cej.2015.10.039.
24. Elessawy, N. A., El-Sayed, E. M., Ali, S., Elkady, M. F., Elnouby, M., & Hamad, H. A. (2020a,
October). One-pot green synthesis of magnetic fullerene nanocomposite for adsorption char-
acteristics. Journal of Water Process Engineering, 34, 101047. https://doi.org/10.1016/j.
jwpe.2019.101047.
25. Elessawy, N. A., Elnouby, M., Gouda, M. H., Hamad, H. A., Taha, N. A., Gouda, M., & Mohy
Eldin, M. S. (2020b). Ciprofloxacin removal using magnetic fullerene nanocomposite obtained
from sustainable PET bottle wastes: Adsorption process optimization, kinetics, isotherm,
regeneration and recycling studies. Chemosphere, 239, 124728. https://doi.org/10.1016/j.
chemosphere.2019.124728.
26. Liu, Y., Peng, Y., An, B., Li, L., & Liu, Y. (2020). Effect of molecular structure on the
adsorption affinity of sulfonamides onto CNTs: Batch experiments and DFT calculations.
Chemosphere, 246, 125778. https://doi.org/10.1016/j.chemosphere.2019.125778.
27. Wang, F., Ma, S., Si, Y., Dong, L., Wang, X., Yao, J., Chen, H., Yi, Z., Yao, W., & Xing,
B. (2017). Interaction mechanisms of antibiotic sulfamethoxazole with various graphene-­
based materials and multiwall carbon nanotubes and the effect of humic acid in water. Carbon,
114(2017), 671–678. https://doi.org/10.1016/j.carbon.2016.12.080.
360 N. Sasirekha and Y.-W. Chen

28. Gao, Y., Li, Y., Zhang, L., Huang, H., Hu, J., Shah, S. M., & Su, X. (2012). Adsorption and
removal of tetracycline antibiotics from aqueous solution by graphene oxide. Journal of
Colloid and Interface Science, 368(1), 540–546. https://doi.org/10.1016/j.jcis.2011.11.015.
29. He, J., Ni, F., Cui, A., Chen, X., Deng, S., Shen, F., Huang, C., Yang, G., Song, C., Zhang, J.,
Tian, D., Long, L., Zhu, Y., & Luo, L. (2020). New insight into adsorption and co-adsorption of
arsenic and tetracycline using a Y-immobilized graphene oxide-alginate hydrogel: Adsorption
behaviours and mechanisms. Science of the Total Environment, 701, 134363. https://doi.
org/10.1016/j.scitotenv.2019.134363.
30. Li, Z., Qi, M., Tu, C., Wang, W., Chen, J., & Wang, A. J. (2017). Highly efficient removal
of chlorotetracycline from aqueous solution using graphene oxide/TiO 2 composite:
Properties and mechanism. Applied Surface Science, 425, 765–775. https://doi.org/10.1016/j.
apsusc.2017.07.027.
31. Ncibi, M. C., & Sillanpää, M. (2015). Optimized removal of antibiotic drugs from aqueous
solutions using single, double and multi-walled carbon nanotubes. Journal of Hazardous
Materials, 298, 102–110. https://doi.org/10.1016/j.jhazmat.2015.05.025.
32. Carrales-Alvarado, D. H., Leyva-Ramos, R., Rodríguez-Ramos, I., Mendoza-Mendoza, E., &
Moral-Rodríguez, A. E. (2020). Adsorption capacity of different types of carbon nanotubes
towards metronidazole and dimetridazole antibiotics from aqueous solutions: Effect of mor-
phology and surface chemistry. Environmental Science and Pollution Research. https://doi.
org/10.1007/s11356-­020-­08110-­x.
33. Kim, H., Hwang, Y. S., & Sharma, V. K. (2014). Adsorption of antibiotics and iopromide onto
single-walled and multi-walled carbon nanotubes. Chemical Engineering Journal, 255, 23–27.
https://doi.org/10.1016/j.cej.2014.06.035.
34. Sun, Y., Chen, M., Liu, H., Zhu, Y., Wang, D., & Yan, M. (2020, April). Adsorptive removal of
dye and antibiotic from water with functionalized zirconium-based metal organic framework
and graphene oxide composite nanomaterial Uio-66-(OH)2/GO. Applied Surface Science,
525, 146614. https://doi.org/10.1016/j.apsusc.2020.146614.
35. Mojiri, A., Zhou, J.  L., Robinson, B., Ohashi, A., Ozaki, N., Kindaichi, T., Farraji, H., &
Vakili, M. (2020). Pesticides in aquatic environments and their removal by adsorption meth-
ods. Chemosphere, 253, 126646. https://doi.org/10.1016/j.chemosphere.2020.126646.
36. Lima, J.  D. M., Gomes, D.  S., Frazão, N.  F., Soares, D.  J. B., & Sarmento, R.  G. (2020).
Glyphosate adsorption on C6 0 fullerene in aqueous medium for water reservoir depollution.
Journal of Molecular Modeling, 26(5). https://doi.org/10.1007/s00894-­020-­04366-­9.
37. Wang, H., Hu, B., Gao, Z., Zhang, F., & Wang, J. (2020). Emerging role of graphene oxide
as sorbent for pesticides adsorption: Experimental observations analyzed by molecular mod-
eling. Journal of Materials Science and Technology, 63, 192–202. https://doi.org/10.1016/j.
jmst.2020.02.033.
38. Mandeep, Gulati, A., & Kakkar, R. (2020). DFT study of adsorption of glyphosate pesticide
on Pt-Cu decorated pyridine-like nitrogen-doped graphene. Journal of Nanoparticle Research,
22(1). https://doi.org/10.1007/s11051-­019-­4730-­z.
39. Marin, P., Bergamasco, R., Nivaldo, A., Roberto, P., & Hamoudi, S. (2019). Synthesis and
characterization of graphene oxide functionalized with MnFe 2 O 4 and supported on acti-
vated carbon for glyphosate adsorption in fixed bed column. Process Safety and Environmental
Protection, 123, 59–71. https://doi.org/10.1016/j.psep.2018.12.027.
40. Vendrame, L. F. O., Zuchetto, T., Fagan, S. B., & Zanella, I. (2019, May). Nanofilter based
on functionalized carbon nanostructures for the adsorption of pentachlorophenol mol-
ecules. Computational and Theoretical Chemistry, 1165, 112561. https://doi.org/10.1016/j.
comptc.2019.112561.
41. Zhang, W., Ruan, G., Li, X., Jiang, X., Huang, Y., Du, F., & Li, J. (2019). Novel porous carbon
composites derived from a graphene-modified high-internal-phase emulsion for highly effi-
cient separation and enrichment of triazine herbicides. Analytica Chimica Acta, 1071, 17–24.
https://doi.org/10.1016/j.aca.2019.04.041.
13  Carbonaceous Nanomaterials for Environmental Remediation 361

42. Ali, I., Alharbi, O. M. L., ALOthman, Z. A., Al-Mohaimeed, A. M., & Alwarthan, A. (2019).
Modeling of fenuron pesticide adsorption on CNTs for mechanistic insight and removal in
water. Environmental Research, 170, 389–397. https://doi.org/10.1016/j.envres.2018.12.066.
43. Al-Shaalan, N. H., Ali, I., ALOthman, Z. A., Al-Wahaibi, L. H., & Alabdulmonem, H. (2019).
High performance removal and simulation studies of diuron pesticide in water on MWCNTs.
Journal of Molecular Liquids, 289, 111039. https://doi.org/10.1016/j.molliq.2019.111039.
44. Dehghani, M. H., Kamalian, S., Shayeghi, M., Yousefi, M., Heidarinejad, Z., Agarwal, S., &
Gupta, V. K. (2019). High-performance removal of diazinon pesticide from water using multi-­
walled carbon nanotubes. Microchemical Journal, 145, 486–491. https://doi.org/10.1016/j.
microc.2018.10.053.
45. Duman, O., Özcan, C., Gürkan Polat, T., & Tunç, S. (2019). Carbon nanotube-based magnetic
and non-magnetic adsorbents for the high-efficiency removal of diquat dibromide herbicide
from water: OMWCNT, OMWCNT-Fe3O4 and OMWCNT-Κ-carrageenan-Fe3O4 nanocom-
posites. Environmental Pollution, 244, 723–732. https://doi.org/10.1016/j.envpol.2018.10.071.
46. Liu, C., Wang, P., Liu, X., Yi, X., Zhou, Z., & Liu, D. (2019a). Multifunctional β-cyclodextrin
MOF-derived porous carbon as efficient herbicides adsorbent and potassium fertilizer.
ACS Sustainable Chemistry & Engineering, 7(17), 14479–14489. https://doi.org/10.1021/
acssuschemeng.9b01911.
47. Mashkoor, F., Nasar, A., & Inamuddin. (2020). Carbon nanotube-based adsorbents for the
removal of dyes from waters: A review. Environmental Chemistry Letters, 18(3), 605–629.
https://doi.org/10.1007/s10311-­020-­00970-­6.
48. Rajabi, M., Mahanpoor, K., & Moradi, O. (2017). Removal of dye molecules from aqueous
solution by carbon nanotubes and carbon nanotube functional groups: Critical review. RSC
Advances, 7, 47083–47090. https://doi.org/10.1039/c7ra09377b.
49. Ferreira, G. M. D., Ferreira, G. M. D., Hespanhol, M. C., de Paula Rezende, J., dos Santos
Pires, A. C., Gurgel, L. V. A., & da Silva, L. H. M. (2017). Adsorption of red azo dyes on multi-­
walled carbon nanotubes and activated carbon: A thermodynamic study. Colloids and Surfaces
A: Physicochemical and Engineering Aspects, 529, 531–540. https://doi.org/10.1016/j.
colsurfa.2017.06.021.
50. Ge, Y. L., Zhang, Y. F., Yang, Y., Xie, S., Liu, Y., Maruyama, T., Deng, Z. Y., & Zhao, X. (2019,
April). Enhanced adsorption and catalytic degradation of organic dyes by nanometer iron oxide
anchored to single-wall carbon nanotubes. Applied Surface Science, 488, 813–826. https://doi.
org/10.1016/j.apsusc.2019.05.221.
51. Molla, A., Li, Y., Mandal, B., Kang, S. G., Hur, S. H., & Chung, J. S. (2019). Selective adsorp-
tion of organic dyes on graphene oxide: Theoretical and experimental analysis. Applied
Surface Science, 464, 170–177. https://doi.org/10.1016/j.apsusc.2018.09.056.
52. Chang, S., Zhang, Q., Lu, Y., Wu, S., & Wang, W. (2020). High-efficiency and selective adsorp-
tion of organic pollutants by magnetic CoFe2O4/graphene oxide adsorbents: Experimental and
molecular dynamics simulation study. Separation and Purification Technology, 238, 116400.
https://doi.org/10.1016/j.seppur.2019.116400.
53. Ghosh, S., Ojha, P. K., & Roy, K. (2019). Exploring QSPR modeling for adsorption of hazard-
ous synthetic organic chemicals (SOCs) by SWCNTs. Chemosphere, 228, 545–555. https://
doi.org/10.1016/j.chemosphere.2019.04.124.
54. Roy, J., Ghosh, S., Ojha, P. K., & Roy, K. (2019). Predictive quantitative structure-property
relationship (QSPR) modeling for adsorption of organic pollutants by carbon nanotubes
(CNTs). Environmental Science: Nano, 6(1), 224–247. https://doi.org/10.1039/c8en01059e.
55. Ma, L., Li, K., Wang, C., Liu, B., Peng, H., Mei, Y., & Ning, P. (2020b). Enhanced

adsorption of hydrophobic organic contaminants by high surface area porous graphene.
Environmental Science and Pollution Research, 27(7), 7309–7317. https://doi.org/10.1007/
s11356-­019-­07439-­2.
56. Jeong, Y., Cui, M., Choi, J., Lee, Y., Kim, J., Son, Y., & Khim, J. (2020). Development of mod-
ified mesoporous carbon (CMK-3) for improved adsorption of bisphenol-A. Chemosphere,
238, 124559. https://doi.org/10.1016/j.chemosphere.2019.124559.
362 N. Sasirekha and Y.-W. Chen

57. Fiyadh, S. S., AlSaadi, M. A., Jaafar, W. Z., AlOmar, M. K., Fayaed, S. S., Mohd, N. S., Hin,
L. S., & El-Shafie, A. (2019). Review on heavy metal adsorption processes by carbon nanotubes.
Journal of Cleaner Production, 230, 783–793. https://doi.org/10.1016/j.jclepro.2019.05.154.
58. Ouni, L., Ramazani, A., & Taghavi Fardood, S. (2019). An overview of carbon nanotubes role
in heavy metals removal from wastewater. Frontiers of Chemical Science and Engineering,
13(2), 274–295. https://doi.org/10.1007/s11705-­018-­1765-­0.
59. Yang, X., Wan, Y., Zheng, Y., He, F., Yu, Z., Huang, J., Wang, H., Ok, Y. S., Jiang, Y., & Gao,
B. (2019, February). Surface functional groups of carbon-based adsorbents and their roles in
the removal of heavy metals from aqueous solutions: A critical review. Chemical Engineering
Journal, 366, 608–621. https://doi.org/10.1016/j.cej.2019.02.119.
60. Singha Deb, A. K., Dhume, N., Dasgupta, K., Ali, S. M., Shenoy, K. T., & Mohan, S. (2019).
Sulphur Ligand Functionalized Carbon Nanotubes for Removal of mercury from waste water–
experimental and density functional theoretical study. Separation Science and Technology
(Philadelphia), 54(10), 1573–1587. https://doi.org/10.1080/01496395.2018.1529044.
61. Ghorbani, M., Seyedin, O., & Aghamohammadhassan, M. (2020, October). Adsorptive

removal of lead (II) ion from water and wastewater media using carbon-based nanomaterials
as unique sorbents: A review. Journal of Environmental Management, 254, 109814. https://doi.
org/10.1016/j.jenvman.2019.109814.
62. Verma, B., & Balomajumder, C. (2020). Surface modification of one-dimensional Carbon
Nanotubes: A review for the management of heavy metals in wastewater. Environmental
Technology and Innovation, 17, 100596. https://doi.org/10.1016/j.eti.2019.100596.
63. Hayati, B., Maleki, A., Najafi, F., Daraei, H., Gharibi, F., & McKay, G. (2017). Super high
removal capacities of heavy metals (Pb2+ and Cu2+) using CNT dendrimer. Journal of
Hazardous Materials, 336, 146–157. https://doi.org/10.1016/j.jhazmat.2017.02.059.
64. Liang, J., Liu, J., Yuan, X., Dong, H., Zeng, G., Wu, H., Wang, H., Liu, J., Hua, S., Zhang,
S., Yu, Z., He, X., & He, Y. (2015). Facile synthesis of alumina-decorated multi-walled car-
bon nanotubes for simultaneous adsorption of cadmium ion and trichloroethylene. Chemical
Engineering Journal, 273, 101–110. https://doi.org/10.1016/j.cej.2015.03.069.
65. Geng, J., Yin, Y., Liang, Q., Zhu, Z., & Luo, H. (2019). Polyethyleneimine cross-linked gra-
phene oxide for removing hazardous hexavalent chromium: Adsorption performance and
mechanism. Chemical Engineering Journal, 361, 1497–1510. https://doi.org/10.1016/j.
cej.2018.10.141.
66. Elgengehi, S. M., El-Taher, S., Ibrahim, M. A. A., Desmarais, J. K., & El-Kelany, K. E. (2020).
Graphene and graphene oxide as adsorbents for cadmium and lead heavy metals: A theo-
retical investigation. Applied Surface Science, 507, 145038. https://doi.org/10.1016/j.
apsusc.2019.145038.
67. Yao, M., Wang, Z., Liu, Y., Yang, G., & Chen, J. (2019, February). Preparation of dialde-
hyde cellulose graftead graphene oxide composite and its adsorption behavior for heavy met-
als from aqueous solution. Carbohydrate Polymers, 212, 345–351. https://doi.org/10.1016/j.
carbpol.2019.02.052.
68. Kong, Q., Preis, S., Li, L., Luo, P., Hu, Y., & Wei, C. (2020). Graphene oxide-terminated
hyperbranched amino polymer-carboxymethyl cellulose ternary nanocomposite for effi-
cient removal of heavy metals from aqueous solutions. International Journal of Biological
Macromolecules, 149, 581–592. https://doi.org/10.1016/j.ijbiomac.2020.01.185.
69. Ahmad, S. Z. N., Wan Salleh, W. N., Ismail, A. F., Yusof, N., Mohd Yusop, M. Z., & Aziz,
F. (2020). Adsorptive removal of heavy metal ions using graphene-based nanomaterials:
Toxicity, roles of functional groups and mechanisms. Chemosphere, 248, 126008. https://doi.
org/10.1016/j.chemosphere.2020.126008.
70. Abdelsalam, H., Teleb, N. H., Yahia, I. S., Zahran, H. Y., Elhaes, H., & Ibrahim, M. A. (2019).
Graphene-based adsorbents for the removal of toxic organic pollutants: A review. Journal of
Physics and Chemistry of Solids, 130, 32–40. https://doi.org/10.1016/j.jpcs.2019.02.014.
13  Carbonaceous Nanomaterials for Environmental Remediation 363

71. Liu, C., Tang, Y., Huo, P., & Chen, F. (2019b). Novel AgCl/CNTs/g-C3N4 nanocomposite
with high photocatalytic and antibacterial activity. Materials Letters, 257, 126708. https://doi.
org/10.1016/j.matlet.2019.126708.
72. Samsudin, M. F. R., Bacho, N., Sufian, S., & Ng, Y. H. (2019). Photocatalytic degradation of
phenol wastewater over Z-scheme g-C 3 N 4/CNT/BiVO 4 heterostructure photocatalyst under
solar light irradiation. Journal of Molecular Liquids, 277, 977–988. https://doi.org/10.1016/j.
molliq.2018.10.160.
73. Nuengmatcha, P., Porrawatkul, P., Chanthai, S., Sricharoen, P., & Limchoowong, N. (2019).
Enhanced photocatalytic degradation of methylene blue using Fe2O3/graphene/CuO nano-
composites under visible light. Journal of Environmental Chemical Engineering, 7(6), 103438.
https://doi.org/10.1016/j.jece.2019.103438.
74. Bai, X., Wang, L., Wang, Y., Yao, W., & Zhu, Y. (2014). Enhanced oxidation ability of g-C3N4
photocatalyst via C60 modification. Applied Catalysis B: Environmental, 152–153(1),
262–270. https://doi.org/10.1016/j.apcatb.2014.01.046.
75. Pan, Y., Liu, X., Zhang, W., Liu, Z., Zeng, G., Shao, B., Liang, Q., He, Q., Yuan, X., Huang,
D., & Chen, M. (2020). Advances in photocatalysis based on fullerene C60 and its derivatives:
Properties, mechanism, synthesis, and applications. Applied Catalysis B: Environmental, 265,
118579. https://doi.org/10.1016/j.apcatb.2019.118579.
76. Ramanathan, A.  A., Aqra, M.  W., & Al-Rawajfeh, A.  E. (2018). Recent advances in 2D
nanopores for desalination. Environmental Chemistry Letters, 16(4), 1217–1231. https://doi.
org/10.1007/s10311-­018-­0745-­4.
77. Zhang, J., Chen, C., Pan, J., Zhang, L., Liang, L., Kong, Z., Wang, X., Zhang, W., & Shen,
J. W. (2020a). Atomistic insights into the separation mechanism of multilayer graphene mem-
branes for water desalination. Physical Chemistry Chemical Physics, 22(14), 7224–7233.
https://doi.org/10.1039/d0cp00071j.
78. Ang, E. Y. M., Toh, W., Yeo, J., Lin, R., Liu, Z., Geethalakshmi, K. R., & Ng, T. Y. (2020). A
review on low dimensional carbon desalination and gas separation membrane designs. Journal
of Membrane Science, 598, 117785. https://doi.org/10.1016/j.memsci.2019.117785.
79. Foroutan, M., Naeini, V. F., & Ebrahimi, M. (2019, February). Carbon nanotubes encapsulat-
ing fullerene as water nano-channels with distinctive selectivity: Molecular dynamics simula-
tion. Applied Surface Science, 489, 198–209. https://doi.org/10.1016/j.apsusc.2019.05.229.
80. Mao, Y., Huang, Q., Meng, B., Zhou, K., Liu, G., Gugliuzza, A., Drioli, E., & Jin, W. (2020).
Roughness-enhanced hydrophobic graphene oxide membrane for water desalination via
membrane distillation. Journal of Membrane Science, 611, 118364. https://doi.org/10.1016/j.
memsci.2020.118364.
81. Hung, W. S., An, Q. F., De Guzman, M., Lin, H. Y., Huang, S. H., Liu, W. R., Hu, C. C., Lee,
K. R., & Lai, J. Y. (2014). Pressure-assisted self-assembly technique for fabricating composite
membranes consisting of highly ordered selective laminate layers of amphiphilic graphene
oxide. Carbon, 68, 670–677. https://doi.org/10.1016/j.carbon.2013.11.048.
82. World Health Organization Report, 2020, https://www.who.int/health-topics/air-pollution#
tab=tab_1.
83. Roy, K., Mukherjee, A., Maddela, N.  R., Chakraborty, S., Shen, B., Li, M., Du, D., Peng,
Y., Lu, F., & Garciá Cruzatty, L.  C. (2020). Outlook on the bottleneck of carbon nanotube
in desalination and membrane-based water treatment—A review. Journal of Environmental
Chemical Engineering, 8(1), 103572. https://doi.org/10.1016/j.jece.2019.103572.
84. Kumar, V., Lee, Y. S., Shin, J. W., Kim, K. H., Kukkar, D., & Fai Tsang, Y. (2020, November).
Potential applications of graphene-based nanomaterials as adsorbent for removal of vola-
tile organic compounds. Environment International, 135, 105356. https://doi.org/10.1016/j.
envint.2019.105356.
85. Liu, Y., Zhang, H., Zhang, Z., Jia, X., & An, L. (2019c, May). CO adsorption on Fe-doped
vacancy-defected CNTs—A DFT study. Chemical Physics Letters, 730, 316–320. https://doi.
org/10.1016/j.cplett.2019.06.013.
364 N. Sasirekha and Y.-W. Chen

86. An, L., Jia, X., & Liu, Y. (2019). Adsorption of SO 2 molecules on Fe-doped carbon nano-
tubes: The first principles study. Adsorption, 25(2), 217–224. https://doi.org/10.1007/
s10450-­019-­00026-­4.
87. Darvish Ganji, M., & Kiyani, H. (2019). Molecular simulation of efficient removal of H2S
pollutant by cyclodextrine functionalized CNTs. Scientific Reports, 9(1), 1–14. https://doi.
org/10.1038/s41598-­019-­46816-­2.
88. Zou, W., Gao, B., Ok, Y. S., & Dong, L. (2019). Integrated adsorption and photocatalytic deg-
radation of volatile organic compounds (VOCs) using carbon-based nanocomposites: A criti-
cal review. Chemosphere, 218, 845–859. https://doi.org/10.1016/j.chemosphere.2018.11.175.
89. Zhang, R., Zhang, A., Yang, Y., Cao, Y., Dong, F., & Zhou, Y. (2020b, January). Surface
modification to control the secondary pollution of photocatalytic nitric oxide removal over
monolithic protonated g-C3N4/graphene oxide aerogel. Journal of Hazardous Materials, 397,
122822. https://doi.org/10.1016/j.jhazmat.2020.122822.
90. Zou, W., Li, X., Lai, Z., Zhang, X., Hu, X., & Zhou, Q. (2016). Graphene oxide inhibits antibi-
otic uptake and antibiotic resistance gene propagation. ACS Applied Materials and Interfaces,
8(48), 33165–33174. https://doi.org/10.1021/acsami.6b09981.
91. Chen, M., Sun, Y., Liang, J., Zeng, G., Li, Z., Tang, L., Zhu, Y., Jiang, D., & Song, B. (2019,
February). Understanding the influence of carbon nanomaterials on microbial communities.
Environment International, 126, 690–698. https://doi.org/10.1016/j.envint.2019.02.005.
92. Verma, S. K., Das, A. K., Gantait, S., Kumar, V., & Gurel, E. (2019). Applications of carbon
nanomaterials in the plant system: A perspective view on the pros and cons. Science of the
Total Environment, 667, 485–499. https://doi.org/10.1016/j.scitotenv.2019.02.409.
93. Peng, Z., Liu, X., Zhang, W., Zeng, Z., Liu, Z., Zhang, C., Liu, Y., Shao, B., Liang, Q., Tang,
W., & Yuan, X. (2020, October). Advances in the application, toxicity and degradation of car-
bon nanomaterials in environment: A review. Environment International, 134, 105298. https://
doi.org/10.1016/j.envint.2019.105298.
Chapter 14
Magnetically Recyclable Photocatalysts
for Degradation of Organic Pollutants
in Aquatic Environment

Ashutosh Kumar and Sushil Kumar Kansal

14.1  Photocatalysis

Photocatalysis—a light-driven catalysis process—employed for degradation of


organic pollutants through activation of a semiconductor material, i.e., photocata-
lyst upon light irradiation. Basically, the degradation of organic pollutants through
photocatalysis process can be achieved in the following five steps (Fig. 14.1) [1, 2]:
1 . Adsorption of organic pollutant on to the photocatalyst’s surface.
2. Absorption of light energy by the photocatalyst, resulting in the photogeneration
of electrons (e−) and holes (h+). The e−–h+ pair can be generated, if the light
energy is more than or equal to the band gap (Eg) of a photocatalyst.
3. Separation of e−–h+ pair.
4. Generation of reactive oxygen species (ROS).
5. Degradation of organic pollutant adsorbed onto the photocatalyst’s surface

through oxidation–reduction reaction carried out by the e−, h+, and ROS.

A. Kumar (*)
School of Energy and Environment, Thapar Institute of Engineering and Technology,
Patiala, Punjab, India
e-mail: ashu.kumar@thapar.edu; ak10907030@gmail.com
S. K. Kansal (*)
Dr. S. S. Bhatnagar University Institute of Chemical Engineering and Technology, Panjab
University, Chandigarh, India
e-mail: sushilkk1@pu.ac.in

© Springer Nature Switzerland AG 2021 365


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_14
366 A. Kumar and S. K. Kansal

Fig. 14.1 Diagram
showing degradation of
organic pollutant in a
typical photocatalysis
process

Fig. 14.2  A typical


diagram showing
generation of ROS onto the
photocatalyst’s surface

14.1.1  Generation of ROS

In fact, upon light irradiation onto the photocatalyst’s surface, the e−–h+ pairs are
generated. By virtue of their opposite charge, the e− and h+ have tendency to recom-
bine each other which is known as recombination process. Furthermore, the effi-
ciency of the photocatalysis process can be enhanced by lowering the recombination
rate of e− and h+, which can provide more time for the generation of ROS. Accordingly,
the low rate of recombination of e− and h+ provides relatively more time for the
generation of ROS through the respective oxidation–reduction reactions with the
reductive e− and oxidative h+, which ultimately results in relatively more number of
ROS available for the pollutant degradation through the photocatalysis process.
A typical diagram showing generation of ROS onto the photocatalyst’s surface is
shown in Fig. 14.2. In a typical photocatalysis process, the oxidative photogene-
rated h+ present in the valence band (VB) reacts with the H2O present in the aquatic
environment to form the hydroxyl radical (•OH). Similarly, the reductive photogen-
erated e− present in the conduction band (CB) reacts with the O2 present in the
aquatic environment to form the superoxide radical (•O2). Due to further proton-
ation of the •O2, it can also form the hydrogen peroxide (H2O2) which can subse-
quently generate the •OH. It is to be noted that the •OH is the most powerful and
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 367

non–selective ROS in the photocatalytic degradation process which has ability to


degrade almost all types of organic pollutants [2, 3].

14.1.2  Challenges in Photocatalysis

Ideally, photocatalysis process leads to mineralization of organic pollutants into


CO2 and H2O, which are relatively less harmful products than their reactants, i.e.,
organic pollutants. However, the basic challenges of the photocatalysis process limit
its application which can be due to (a) the activity of the photocatalyst in relatively
higher energy or lower wavelength of light present in solar spectrum (e.g., UV)
only; (b) the high recombination rate of e−–h+ pair, providing relatively low resi-
dence time for the generation ROS in the solution which in turn produces relatively
lower number of ROS; and (c) dissipation of the captured light energy though
recombination of e−–h+ pair [2, 4].
In order to understand the limited activity of the photocatalyst in the lower wave-
length of light, the solar spectrum is shown in Fig.  14.3. As shown in Fig.  14.3,
when one moves from the UV region toward the visible light or infrared (IR) region,
the light energy has relatively lower energy but higher wavelength. It depicts that
activity in lower wavelength region requires higher energy and thus the higher cost
of photocatalysis process from its application point of view. Moreover, the UV light
shares only 2–3% of the solar radiation, while the visible light accounts to ~45% of
the total solar radiation [6, 7]. Therefore, designing a photocatalyst which can be
activated by (a) the major share of solar spectrum, i.e., visible light, or (b) by most
of the part of solar radiation, i.e., by UV, visible light, and IR together, can over-
come one of the major limitations of the photocatalysis process [8, 9]. The high rate
of recombination of e−–h+ pair is another major limitation of the photocatalysis
process. Briefly, the recombination can be overcome by adopting any of the

Fig. 14.3  Solar spectral


irradiance at air mass 1.5
(AM 1.5) collected from
American Society for
Testing and Materials
(ASTM). Standard: ASTM
G-173-03. Reproduced
with permission from
Elsevier [5]
368 A. Kumar and S. K. Kansal

following strategies: (a) by developing a heterojunction [2, 10], (b) by regulating a


defect or vacancy in the photocatalyst [11, 12], or (c) by doping a photocatalyst with
metals or non-metals [7, 13, 14]. It is to be noted that the challenges of the photoca-
talysis process is not the major focus of this chapter, and thus only important limita-
tions vis-à-vis challenges are discussed.

14.2  Organic Pollutants in Aquatic Environment

Aquatic environment, i.e., water and wastewater, contains various types of organic
pollutants. Any undesirable chemical substance of organic nature which can poten-
tially affect the human health are categorized as an organic pollutant. From the lit-
erature, it is observed that the researchers have targeted various groups of organic
pollutants for degradation, such as dyes [15, 16], endocrine disrupting chemicals
(EDCs) [17, 18], pharmaceuticals and personal care products (PPCPs) [19–21], per-
sistent organic pollutants (POPs) [22, 23], poly-and per-fluoroalkyl substances
(PFAS) [24, 25], emerging chemicals [26, 27], etc.
PFAS are a group of anthropogenic substances with amphiphilic properties
which are used in our day to day life as fire suppressants, stain and water repellants.
After their use, the PFAS find their ways into the aquatic environment and thereafter
to the human being and animals, after consuming such water. The PFAS can cause
adverse immunological effects, reproductive disorders, liver ailments in animals,
and immuno-toxicological and neuro-developmental issues in children [24, 25].
PPCPs represent a group of chemicals containing pharmaceuticals, viz., analgesics,
antibiotics, anti-inflammatory, anti-epileptic drugs, etc., and personal care products,
viz., insect repellants, synthetic musks, UV filters, etc. The adverse effects of these
chemicals on aquatic lives include birth defects, endocrine disruptions, oxidation
stress, post-embryonic developmental issues, and reproductive disorder [7, 20, 21].
Dyes are another class of visibly recognized harmful pollutants which are dis-
charged into water bodies by various industries, viz., food, hair dyeing, leather tan-
ning, paper and pulp, photographic, printing, textile, etc. The dyes and their toxic
metabolites are known to cause bladder cancer in human being [14, 28]. EDCs are
a group of exogenous chemical substances which can alter the function of endocrine
system and consequently causes adverse effects in an organism or its progeny. These
chemicals are brominated flame retardants, dioxins, bisphenol A, organic solvents,
parabens, phthalates, pesticides, polycyclic aromatic hydrocarbons (PAHs), heavy
metals, some naturally occurring phytoestrogens, etc. The EDCs can potentially
cause developmental abnormalities and reproductive issues in aquatic lives, birds,
and animals [17, 29]. POPs are a group of chemicals which remain in the environ-
ment for longer period, and hence they can bioaccumulate and transported into the
food chain. The POPs include chemicals, such as herbicides, insecticides, fungi-
cides, pesticides, etc., which are potentially harmful to the environment and human
health [3, 23]. Considering adverse effects of these organic pollutants on human
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 369

being and aquatic environment, their effective treatment becomes a quintessen-


tial task.
Wastewater treatment plants (WWTPs) are considered as the largest sink of these
organic pollutants in aquatic environment. In fact, due to inefficient treatment of
these organic pollutants by WWTPs, they find their ways into receiving water bod-
ies and thereafter they become part of the aquatic environment [2]. Most of these
pollutants are not efficiently treated by the conventional water and wastewater treat-
ment processes, and thus they again become part of the aquatic environment. Thus,
due to their inefficient treatment and regular discharge in the aquatic environment,
their concentration is expected to increase continuously and so is their scale of
related acute and chronic toxicity to the human being and aquatic lives. Alternatively,
photocatalysis, being the effective treatment process for degradation of these organic
pollutants, has invited immense attention from the researchers across the globe. The
key features of photocatalytic degradation process, its mechanism of generation of
ROS, and its basic challenges in degradation of organic pollutants are already dis-
cussed in Sect. 14.1 of this chapter.

14.3  Magnetically Recyclable Photocatalysts

Recycling process consists of separation, regeneration, and reuse of a photocatalyst.


The successful recycling of a photocatalyst can potentially reduce the total cost of
the photocatalysis process by repetitive use of the same photocatalytic material.
Generally, the repetitive use of photocatalytic material can be performed by adopt-
ing one of the following two strategies: (a) by immobilizing the photocatalyst on a
stable support, such as clay, glass, silica, zeolite, etc. [30–32], or (b) by suspending
the photocatalysts in a suspension solution and its recovery after the photocatalytic
degradation process [33–35]. However, the use of former strategy can potentially
reduce the overall surface area of the photocatalyst which can in turn affect the
photocatalytic performance. Thus, the application of photocatalysts in a suspension
solution and their recovery after the photocatalytic degradation process has become
a relatively more popular strategy than immobilizing the photocatalyst on a stable
support.
In recycling process, the separation of photocatalysts from the suspension solu-
tion is usually conducted by adopting the ways, such as (a) centrifugation or (b)
magnetic separation. In laboratory, the centrifugation is performed by utilizing a
centrifuge wherein the photocatalysts and solvents are separated by applying the
mechanical force (centrifugal force). However, the magnetic separation of photo-
catalysts in the laboratory is performed either by employing a handheld laboratory
magnet or by employing an electromagnetic separation system [2, 36]. Due to the
easy handling of magnetic separation process, it has attracted immense attention
among the researchers. Accordingly, the magnetic photocatalysts were developed in
various ways by adopting different strategies. On the basis of the role of magnetic
material in a photocatalyst or composite photocatalyst, it can be divided into two
370 A. Kumar and S. K. Kansal

broad categories: (a) magnetic material as a photocatalyst, (b) magnetic material as


a non-photocatalyst [37, 38].
At large scale, the magnetic separation of magnetic photocatalysts is considered
to be a challenging task. In industries, the magnetic separation of magnetic nanopar-
ticles can be performed by applying two types of magnetic separators, such as high
gradient magnetic separator (HGMS) and open gradient magnetic separator
(OGMS). In HGMS, the separator column provides a high magnetic field gradient
which enables the efficient separation of magnetic nanoparticles, and thus the outlet
of the separator provides a clean solution. It should be noted that the gradient inside
the separator is generated by the ferromagnetic matrix. However, clogging of the
filter by separated magnetic material is the major issue faced by such separators [39,
40]. Alternatively, in OGMS, the magnetic system is arranged in a convenient way
around the wall of the separator in switch mode. When the magnetic system is
switched on, the magnetic photocatalysts can be retained around the wall, and the
separated solution can be obtained at the outlet [40, 41]. The advantage of the
OGMS over HGMS is that it has high separation efficacy while having lower gradi-
ent of the magnetic field. Therefore, at large scale, magnetic separators can be
selected by considering the pros and cons of these two magnetic separators.
After photocatalytic reaction, regeneration of photocatalysts is considered to be
a major challenge. However, regeneration of magnetic photocatalysts after their
magnetic separation can potentially reduce the total cost of the photocatalytic deg-
radation process by reusing the same material, i.e., photocatalyst for multiple cycles.
Recently, Kumar et al. [2] have reviewed the various regeneration methods applied
to regenerate the photocatalysts. After reviewing the various regeneration methods,
it was found that the regeneration was achieved either by washing with regeneration
solution, such as de-ionized (DI) water, organic solvents, H2O2, or by treatment with
gases, such as O3. Among these regeneration methods, it was found that washing
with DI water as well as ethanol followed by drying overnight at 60°C is the most
effective and easy to handle regeneration method. Therefore, in recycling process,
separation and regeneration methods play a crucial role.

14.3.1  Magnetic Material as Photocatalyst

The magnetic material used as a photocatalyst in a composite photocatalyst pos-


sesses two functions: (a) involves in photocatalytic reactions and (b) enables mag-
netic separation of the whole composite photocatalyst due to its magnetic property.
From the literature, it is evident that the magnetic material utilized for the photoca-
talysis process can play the various roles, viz., as a (a) catalyst, [42, 43] (b) co-­
catalyst [12], or (c) dopant [44]. It is to be noted that the magnetic materials utilized
in photocatalytic role are mostly iron-based materials, such as FeO, Fe2O3, Fe3O4,
ferrites, etc. Among the various phases of iron oxides, FeO, Fe2O3, and Fe3O4 (mag-
netite) are the most frequently used magnetic materials in the magnetic photocataly-
sis process. The Fe (III) oxide exists in various forms, such as α–Fe2O3, β–Fe2O3,
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 371

γ–Fe2O3, and ε–Fe2O3. Among these phases of Fe (III) oxide, α–Fe2O3 (hematite),
and γ–Fe2O3 (maghemite) are the more stable forms, and thus they are frequently
used in the photocatalysis process. Moreover, ferrite (MFe2O4) is a transition metal
oxide-based spinal structure in which M represents Co, Fe, Mn, Ni, Zn, etc. [45].
The Fe2O3 possesses a band gap of ~2.3 eV which makes it a visible light active
photocatalyst. It is to be noted that the α–Fe2O3 is found to be stable in the aqueous
solution of pH > 3 which makes it a suitable candidate for the photocatalytic degra-
dation of dyes in aqueous solution [42]. Morphology manipulation is another
approach which can further enhance the photocatalytic efficiency of a photocata-
lyst. In order to improve the photocatalytic efficiency of the α–Fe2O3 photocatalyst,
Zhang et al. [43] developed various morphology of α–Fe2O3, such as nanoparticles,
mesoporous nanorods, and microplates. The scanning electron microscope (SEM)
images of the various morphology of α–Fe2O3 are shown in Fig. 14.4 which are the
as-synthesized nanoparticles (Fig. 14.4a), mesoporous nanorods (Fig. 14.4b), and
microplates (Fig.  14.4c) of α–Fe2O3. Accordingly, the visible light-driven photo-
catalytic activity of the various morphology of α–Fe2O3 was tested for degradation
of methylene blue (MB) dye (Fig. 14.5). As shown in Fig. 14.5, among the various
morphology of α–Fe2O3 photocatalysts, mesoporous α–Fe2O3 nanorods have shown
relatively better photocatalytic performance than the other morphologies for degra-
dation of 100 mL MB of concentration 10−5 M. Moreover, in the presence of H2O2,
the activity of the mesoporous α–Fe2O3 nanorods was found to be further improved
than the mesoporous α–Fe2O3 nanorods alone. After photocatalytic degradation of
MB, the α–Fe2O3 nanorods were separated, regenerated, and employed again for the
degradation of MB under visible light. From the recycling experiments, it was found
that over 82% of photocatalytic efficiency of the mesoporous α–Fe2O3 nanorods
was maintained after 5 cycles. However, the photocatalytic activity of the α–Fe2O3
was still found to be restricted by the common challenges of photocatalysis, i.e., the
high rate of e−–h+ recombination. Therefore, efforts were made to combine such
photocatalyst with another photocatalyst which were expected to lower the rate of
recombination of e−–h+ pair, resulting in improved photocatalytic performance.
The magnetic photocatalysts, viz., Fe2O3 and Fe3O4, have visible light absorption
ability. However, their main limitation as a photocatalyst is the high recombination
rate of charge carriers. Accordingly, efforts were made to combine such magnetic
photocatalysts with other semiconductor materials with superior photocatalytic
ability, such as BiOBr, BiVO4, g–C3N4, graphene oxide, SnO2, TiO2, ZnO, etc. [2,
45, 46]. In such cases, the magnetic photocatalyst acts as a co–catalyst in a compos-
ite photocatalyst, resulting in the low recombination rate of charge carriers and gen-
eration of more number of ROS which are ultimately responsible for photocatalytic
degradation of organic pollutants. For instance, γ–Fe2O3 nanosheets, a 2D gra-
phene–like nanostructure, have excellent electrochemical and optical properties.
However, in photocatalysis process, the γ–Fe2O3 nanosheets suffer from the limita-
tions, such as short life time of photogenerated charge carriers (~10 ps) and their
low diffusion length (2–4  nm) [47]. The TiO2 has excellent corrosion resistance,
chemical inertness, low toxicity, and high redox ability. However, due to its high
372 A. Kumar and S. K. Kansal

Fig. 14.4  SEM image of


α–Fe2O3 (a) nanoparticles,
(b) mesoporous nanorods,
and (c) microplates.
Reproduced with
permission from
Elsevier [43]
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 373

Fig. 14.5 Photodegradation
efficiencies of MB solutions
on various α–Fe2O3
photocatalysts. Reproduced
with permission from
Elsevier [43]

Fig. 14.6 (a) UV–vis spectra, and (b) Magnetization hysteresis loops of α–Fe2O3/b–TiO2 and γ–
Fe2O3/b–TiO2 heterojunctions. Reproduced with permission from Elsevier [12]

band gap (~3.2  eV), it can be activated upon irradiation of UV light only [2].
Alternatively, the black TiO2 (b–TiO2) can show an extended light absorption of the
solar spectrum. The b–TiO2 is known to absorb the light up to the near infrared
(NIR) region, resulting in a higher photogeneration of charge carriers, and thus an
improved photocatalytic performance [48].
Considering the limitations of the γ–Fe2O3 nanosheets and superiority of the b–
TiO2 in the photocatalysis process, Ren et al. [12] developed a γ–Fe2O3/black TiO2
heterojunction photocatalyst for the solar light-driven photocatalytic degradation of
a PPCP, i.e., tetracycline. Briefly, the magnetic ultrathin γ–Fe2O3 nanosheets were
hybridized with the mesoporous black TiO2 (b–TiO2) hollow spheres to form a γ–
Fe2O3/b–TiO2 heterojunction photocatalyst, wherein the γ–Fe2O3 acted as a co-­
catalyst. In order to show extended light absorption in the solar spectrum and
improved magnetic recyclability performance of the heterojunction photocatalyst,
Ren et al. [12] synthesized two types of heterojunction photocatalyst with the same
method, i.e., α–Fe2O3/b–TiO2 and γ–Fe2O3/b–TiO2 heterojunction photocatalysts.
The UV–vis absorption spectra and saturation magnetization (Ms) values obtained
from the the vibrating sample magnetometry (VSM) analysis are shown in
Fig.  14.6a, b, respectively. As shown in Fig.  14.6a, the γ–Fe2O3/b–TiO2
374 A. Kumar and S. K. Kansal

heterojunction photocatalyst has shown the improved visible light absorption com-
pared to the α–Fe2O3/b–TiO2 heterojunction photocatalyst. Moreover, the γ–
Fe2O3/b–TiO2 heterojunction photocatalyst has also shown absorption in the NIR
region of the solar spectrum. From Fig. 14.6b, the Ms value of α–Fe2O3/b–TiO2 and
γ–Fe2O3/b–TiO2 heterojunction photocatalysts were found to be 0.68 and 8.17 emu/g,
respectively, which depicts the improved magnetic behavior of the γ–Fe2O3/b–TiO2
heterojunction photocatalyst compared to the α–Fe2O3/b–TiO2 heterojunction pho-
tocatalyst. Thus, by employing the γ–Fe2O3 as a co–catalyst, a γ–Fe2O3/b–TiO2 het-
erojunction photocatalyst with extended light absorption and improved magnetization
value was developed.
The schematic diagram of the band structures and charge transfer mechanism in
the γ–Fe2O3/b–TiO2 heterojunction photocatalyst are shown in Fig. 14.7. From the
migration of the photogenerated charge carriers onto the heterojunction photocata-
lyst’s surface, it is depicted to be a type–II heterojunction photocatalyst, wherein the
h+ generated in the VB of γ–Fe2O3 migrates toward the VB of b–TiO2 and the pho-
togenerated e− in the CB of TiO2 migrates toward the CB of γ–Fe2O3 nanosheets.
Such migration of the charge carriers, i.e., e− and h+ in this type–II heterojunction
photocatalyst, results in relatively low rate of recombination of charge carriers
which ultimately leads to an improved photocatalytic efficiency. Thus, by employ-
ing γ–Fe2O3 as a co-catalyst in a heterojunction, their photocatalytic efficiency can
be ultimately improved along with their attractive feature of magnetic
recyclability.
Doping of photocatalyst is another strategy for improving the photocatalytic effi-
ciency of the semiconductor photocatalyst that can be achieved either by extending
the light absorption ability of the photocatalyst or by lowering the e−–h+ recombina-
tion rate. For instance, TiO2 is a stable ideal semiconductor which is activated only
after irradiation of UV light. The doping in the TiO2 lattice can be achieved by
adopting the three different strategies, which are shown in Fig. 14.8. As shown, the
main rationale behind the various strategies of doping in TiO2 is to decrease its wide
band gap which can be achieved either by a lower shift of conduction band mini-
mum (CBM) (Fig.  14.8a), a higher shift of valence band maximum (VBM)
(Fig. 14.8b), or creating the impurity states (Fig. 14.8c).

Fig. 14.7 Schematic
illustration of the energy
band structure for
γ–Fe2O3/b–TiO2
heterojunctions and the
proposed photogenerated
charge transfer mechanism.
Reproduced with
permission from
Elsevier [12]
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 375

Fig. 14.8  Three schemes of the band gap modifications for visible light sensitization with a lower
shift of CBM (a), a higher shift of VBM (b), and impurity states (c). Reprinted with permission
from [44]. Copyright (2014) American Chemical Society

Fig. 14.9  Comparison of


atomic p levels among
anions. The band gap of
TiO2 is formed between O
2pπ and Ti 3d states.
Reprinted with permission
from Asahi et al. [44].
Copyright (2014)
American Chemical
Society

To understand the feasibility of various strategies of doping in the TiO2 lattice, an


atomic level energy diagram of the TiO2, showing the O 2pπ and Ti d states in the
band gap, is shown in Fig. 14.9. As shown, the CBM of TiO2 consists of the Ti d
states, where doping can be easily achieved by substituting the Ti with 3d transition
metals, viz., Cr, Fe, Mn, V, etc. Contrarily, the VBM of TiO2 contains the O 2pπ
states, where doping can be easily achieved by replacing the anions, containing the
atomic p levels, viz., B, C, F, O, P, N, S, etc. It is to be noted that the doping in the
CBM can be performed mostly by the metals, while the doping in the VBM can be
performed by the non-metals. However, the impurity levels can be created both by
metals and non-metals [2, 44]. Overall, the magnetic material as a photocatalyst is
found to have their wide application in the photocatalysis process, such as (a) cata-
lyst, (b) co-catalyst, or (c) dopant, wherein their application as a co-catalyst is found
to be more promising in the photocatalytic degradation of organic pollutants by
virtue of the fact that it has ability to extend the light absorption in the wide range
of solar spectrum, i.e., UV to NIR region.
376 A. Kumar and S. K. Kansal

14.3.2  Magnetic Material as Non-photocatalyst

Generally, the non-photocatalytic use of magnetic material in a composite photo-


catalyst is to facilitate the magnetic separation of the composite photocatalyst from
the suspension after use. In such cases, the main photocatalytic material is insulated
from the magnetic material by providing a non-conductive reaction barrier. For
instance, Kumar et  al. [49] developed a terephthalic acid functionalized g–C3N4/
TiO2 heterojunction photocatalyst which was separated from the magnetic Fe3O4
(magnetite) nanoparticles using a non-conductive SiO2 layer through formation of a
core–shell Fe3O4@SiO2 structure. The transmission electron microscope (TEM)
image of an unique core–shell Fe3O4@SiO2 nanoparticles can be seen in Fig. 14.10a,
wherein the SiO2 layer of ~5 nm thickness can be seen encapsulating the Fe3O4 core
of thickness ~22 nm containing the Fe3O4 nanoparticles. Similarly, Álvarez et al.
[50] developed a TiO2/SiO2/Fe3O4 photocatalyst for degradation of PPCPs under
UV light irradiation. Figure 14.10b shows the TEM image of the TiO2/SiO2/Fe3O4
photocatalyst where the TiO2 shell can be clearly seen over the non-conductive SiO2
layer. Moreover, the Ms values of the visible light-driven terephthalic acid function-
alized g–C3N4/TiO2/Fe3O4@SiO2 and UV–light–driven TiO2/SiO2/Fe3O4 composite
photocatalysts were found to be 8 and 40  emu/g, respectively. It should also be
noted that the Ms value of 8 emu/g is the sufficient Ms value which can help in the
magnetic recycling of the composite photocatalysts. In such cases, the SiO2 shell
prevents the flow of charge carriers from the main photocatalyst toward the Fe3O4
nanoparticles (also known as the photodissolution effect) which can ultimately
result in the reduction of overall photocatalytic degradation efficiency. Furthermore,
the SiO2 shell also protects the Fe3O4 core nanoparticles from the oxidation which
can help in maintaining its superparamagnetic behavior. Therefore, by utilizing
such strategies, the magnetically recyclable photocatalysts can be developed where
the magnetic material is in non-photocatalytic role.

Fig. 14.10 (a) TEM image


of as-synthesized core–
shell Fe3O4@SiO2
nanoparticle and (b) TEM
image of TiO2/SiO2/Fe3O4
nanoparticles. Reproduced
with permission from
Elsevier [49, 50]
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 377

14.4  Magnetic Recycling: Practical Challenges Ahead

As mentioned in Sect. 14.3, at laboratory scale, the magnetic recycling of the pho-
tocatalysts can be performed either by using a handheld laboratory magnet or by
employing an electromagnetic separation unit. Various researchers have developed
the visible light-driven magnetically recyclable photocatalysts for degradation of
organic pollutants, where magnetic separation of photocatalysts from the suspen-
sion solution was performed by using a handheld laboratory magnet [4, 36, 38].
However, very few researchers have also performed the degradation of organic pol-
lutant in a visible light-driven prototype reactor rather than in a bench scale reactor.
In this regard, after photocatalytic degradation of organic pollutants, i.e., PPCPs,
Kumar et al. [7] and Khan et al. [4] were able to separate the magnetic photocata-
lysts from the suspension solution by employing a ~200 mT electromagnetic mag-
netic separation unit. In fact, the electromagnetic separation unit in these studies
was part of a 5 L visible light-driven prototype reactor system. The schematic dia-
gram (Fig. 14.11a) and the real photographs of the photocatalytic reactor system
(Fig. 14.11b) are shown in Fig. 14.11. As shown in Fig. 14.11b, the main photocata-
lytic reactor unit was made of a 5 L capacity cylindrical acrylic glass unit, encom-
passed with the multiple visible light sources of integrated irradiance 317 W/m2. For
the treatment of large amount of real PPCP-laden wastewater, high capacity of such
photocatalytic reactor system is required. In such real scenario, the up-scaling cost
of the photocatalytic reactor system would be very high due to the higher cost of the
electromagnetic separation unit. Moreover, the installation of an electromagnetic
separation unit is another limitation that is due to its large size which could require
more space. Similarly, the installation of the real magnet in place of a magnetic
separation unit is also not suggested due to the similar reason. Therefore, more
research work is required to scale-up the photocatalytic reaction unit with an inte-
grated magnetic separation unit for degradation of organic pollutants by a magneti-
cally recyclable photocatalyst in real practical application.
Generally, the photocatalytic degradation efficiency of the magnetically recy-
clable photocatalysts in the laboratory is tested for organic pollutants present in DI
water. However, in real practical applications, other than the organic pollutants, the
aqueous environment contains various other chemical substances, such as anions,
cations, degradation by-products, humic acid, natural organic matter, etc., which are
proven to be detrimental to the photocatalysis process [2]. In fact, the substances,
such as humic acid and natural organic matter, either work as scavenger for ROS or
reduce the light absorption on the photocatalyst’s surface by shielding the light irra-
diated in the aqueous solution. Moreover, the anions, cations, or degradation by-
products either deactivate the active sites of the photocatalysts or scavenge the ROS
which results in decreased photocatalytic degradation efficiency. Therefore, it is
required to test the environmental effectiveness of the magnetically recyclable pho-
tocatalysts for degradation of organic pollutants present in the real aqueous
environment.
378 A. Kumar and S. K. Kansal

Wastewater

Visible light source

5 L photocatalytic reactor
Electromagnetic system

Magnetic
P separation
unit

P Treated water
Recovered magnetic
photocatalyst

Fig. 14.11 (a) Prototype photocatalytic reactor integrated with a magnetic separation unit and (b)
photographs of the developed prototype photocatalytic reactor integrated with a magnetic separa-
tion unit. Reproduced with permission from Elsevier [4]

14.5  Conclusions and Future Prospects

Photocatalysis, a light-driven catalysis process, has been inviting immense attention


among the researchers worldwide due to its effectiveness for degradation of organic
pollutants present in water and wastewater. The effectiveness of the process is
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 379

known to be due to its ability to mineralize the organic pollutants into relatively less
harmful CO2 and H2O. However, the basic challenges of the photocatalysis process
limit its application due to (a) the activity of the photocatalyst in relatively higher
energy light only (e.g., UV light of the solar spectrum), (b) the high recombination
rate of e−–h+ pair, resulting in dissipation of the captured light energy, and (c) the
recovery of the photocatalysts after its use.
Recycling process consists of separation, regeneration, and reuse of a photocata-
lyst. The successful recycling of a photocatalyst can potentially reduce the total cost
of the photocatalysis process by repetitive use of the same photocatalytic material.
Due to importance of the recycling process, the strategies, viz., (a) immobilizing the
photocatalyst on a stable support, such as clay, glass, silica, zeolite, etc. or (b) sus-
pending the photocatalyst in a suspension solution and its recovery after the photo-
catalytic degradation process, have been adopted. The later strategy has become a
more popular strategy than the former one, due to the fact that it can provide more
surface area for the photocatalytic reaction. Thus, the development of a magneti-
cally recyclable photocatalyst was found to be a way forward which can effectively
facilitate the separation of photocatalysts from the suspension solution after its use.
Based on the role of magnetic material in a photocatalyst or composite photo-
catalyst, such photocatalysts are divided into two broad categories: (a) magnetic
material as a photocatalyst and (b) magnetic material as a non-photocatalyst. The
magnetic material utilized as a photocatalyst can play roles, such as (a) catalyst, (b)
co-catalyst, or (c) dopant. Among these roles, the application of magnetic material
as a co-catalyst is found to be the more promising role due to its ability to extend the
light absorption in the wide range of solar spectrum (i.e., UV to NIR region). The
magnetic material in a non-photocatalytic role is usually insulated from the photo-
catalytic material by a non-conductive reaction barrier (e.g., SiO2). Moreover, the
photocatalytic degradation of organic pollutants at large scale has certain limitations
which can be overcome by (a) scaling–up the photocatalytic reaction unit with an
integrated magnetic separation unit and (b) ensuring the environmental effective-
ness of the photocatalysts in the real aqueous environment. Therefore, the applica-
tion of magnetically recyclable photocatalysts for real practical purposes requires
more research in these areas in future.

References

1. Chong, M. N., Jin, B., Chow, C. W., & Saint, C. (2010). Recent developments in photocatalytic
water treatment technology: A review. Water Research, 44(10), 2997–3027.
2. Kumar, A., Khan, M., He, J., & Lo, I. M. C. (2020a). Recent developments and challenges in
practical application of visible–light–driven TiO2–based heterojunctions for PPCP degrada-
tion: A critical review. Water Research, 170, 115356.
3. Lee, K. M., Lai, C. W., Ngai, K. S., & Juan, J. C. (2016). Recent developments of zinc oxide
based photocatalyst in water treatment technology: A review. Water Research, 88, 428–448.
380 A. Kumar and S. K. Kansal

4. Khan, M., Fung, C.  S., Kumar, A., & Lo, I.  M. C. (2019). Magnetically separable BiOBr/
Fe3O4@ SiO2 for visible–light–driven photocatalytic degradation of ibuprofen: Mechanistic
investigation and prototype development. Journal of Hazardous Materials, 365, 733–743.
5. Islam, M.  S. (2017). Analytical modeling of organic solar cells including monomolecular
recombination and carrier generation calculated by optical transfer matrix method. Organic
Electronics, 41, 143–156.
6. Kaur, A., Anderson, W. A., Tanvir, S., & Kansal, S. K. (2019). Solar light active silver/iron
oxide/zinc oxide heterostructure for photodegradation of ciprofloxacin, transformation prod-
ucts and antibacterial activity. Journal of Colloid and Interface Science, 557, 236–253.
7. Kumar, A., Khan, M., Fang, L., & Lo, I. M. C. (2019). Visible–light–driven N–TiO2@ SiO2@
Fe3O4 magnetic nanophotocatalysts: Synthesis, characterization, and photocatalytic degrada-
tion of PPCPs. Journal of Hazardous Materials, 370, 108–116.
8. Mohamed, M.  A., Zain, M.  F. M., Minggu, L.  J., Kassim, M.  B., Amin, N.  A. S., Salleh,
W.  N. W., Salehmin, M.  N. I., Nasir, M.  F. M., & Hir, Z.  A. M. (2018). Constructing bio-­
templated 3D porous microtubular C–doped g–C3N4 with tunable band structure and enhanced
charge carrier separation. Applied Catalysis B: Environmental, 236, 265–279.
9. Wang, F., Wu, Y., Wang, Y., Li, J., Jin, X., Zhang, Q., Li, R., Yan, S., Liu, H., Feng, Y., & Liu,
G. (2019). Construction of novel Z–scheme nitrogen–doped carbon dots/{001} TiO2 nanosheet
photocatalysts for broad–spectrum–driven diclofenac degradation: Mechanism insight, prod-
ucts and effects of natural water matrices. Chemical Engineering Journal, 356, 857–868.
10. Lu, Z., Zeng, L., Song, W., Qin, Z., Zeng, D., & Xie, C. (2017). In situ synthesis of C–TiO2/g–
C3N4 heterojunction nanocomposite as highly visible light active photocatalyst originated from
effective interfacial charge transfer. Applied Catalysis B: Environmental, 202, 489–499.
11. Che, H., Liu, L., Che, G., Dong, H., Liu, C., & Li, C. (2019). Control of energy band, layer
structure and vacancy defect of graphitic carbon nitride by intercalated hydrogen bond effect
of NO3− toward improving photocatalytic performance. Chemical Engineering Journal, 357,
209–219.
12. Ren, L., Zhou, W., Sun, B., Li, H., Qiao, P., Xu, Y., Wu, J., Lin, K., & Fu, H. (2019). Defects–
engineering of magnetic γ–Fe2O3 ultrathin nanosheets/mesoporous black TiO2 hollow sphere
heterojunctions for efficient charge separation and the solar–driven photocatalytic mechanism
of tetracycline degradation. Applied Catalysis B: Environmental, 240, 319–328.
13. Jiang, L., Yuan, X., Pan, Y., Liang, J., Zeng, G., Wu, Z., & Wang, H. (2017). Doping of gra-
phitic carbon nitride for photocatalysis: A review. Applied Catalysis B: Environmental, 217,
388–406.
14. Sood, S., Umar, A., Mehta, S. K., & Kansal, S. K. (2015). Highly effective Fe–doped TiO2
nanoparticles photocatalysts for visible–light driven photocatalytic degradation of toxic
organic compounds. Journal of Colloid and Interface Science, 450, 213–223.
15. Ahmed, S., & Ahmad, Z. (2020). Development of hexagonal nanoscale nickel ferrite for the
removal of organic pollutant via photo-fenton type catalytic oxidation process. Environmental
Nanotechnology, Monitoring & Management, 14, 100321. https://doi.org/10.1016/j.
enmm.2020.100321.
16. Ahmed, S., Guo, Y., Li, D., Tang, P., & Feng, Y. (2018). Superb removal capacity of hierarchi-
cally porous magnesium oxide for phosphate and methyl orange. Environmental Science and
Pollution Research, 25, 24907–24916.
17. Pelaez, M., Nolan, N. T., Pillai, S. C., Seery, M. K., Falaras, P., Kontos, A. G., Dunlop, P. S.,
Hamilton, J. W., Byrne, J. A., O’shea, K., & Entezari, M. H. (2012). A review on the visible
light active titanium dioxide photocatalysts for environmental applications. Applied Catalysis
B: Environmental, 125, 331–349.
18. Yu, Y., Huang, Q., Wang, Z., Zhang, K., Tang, C., Cui, J., Feng, J., & Peng, X. (2011).
Occurrence and behavior of pharmaceuticals, steroid hormones, and endocrine–disrupting per-
sonal care products in wastewater and the recipient river water of the Pearl river delta, South
China. Journal of Environmental Monitoring, 13(4), 871–878.
14  Magnetically Recyclable Photocatalysts for Degradation of Organic Pollutants… 381

19. Kaur, A., Umar, A., Anderson, W. A., & Kansal, S. K. (2018). Facile synthesis of CdS/TiO2
nanocomposite and their catalytic activity for ofloxacin degradation under visible illumination.
Journal of Photochemistry and Photobiology A: Chemistry, 360, 34–43.
20. Khan, M., Kumar, A., He, J., & Lo, I. M. C. (2020). Elucidating the predominant role of crystal
disorders in hierarchical photocatalysts governing their charge carrier separation and associ-
ated activity in photocatalytic water treatment. Journal of Colloid and Interface Science, 573,
336–347.
21. Liu, J. L., & Wong, M. H. (2013). Pharmaceuticals and personal care products (PPCPs): A
review on environmental contamination in China. Environment International, 59, 208–224.
22. Duttagupta, S., Mukherjee, A., Bhattacharya, A., & Bhattacharya, J. (2020). Wide exposure of
persistent organic pollutants (POPs) in natural waters and sediments of the densely populated
Western Bengal basin, India. Science of the Total Environment, 717, 137187.
23. Fu, J., Mai, B., Sheng, G., Zhang, G., Wang, X., Xiao, X., Ran, R., Cheng, F., Peng, X., Wang,
Z., & Tang, U. W. (2003). Persistent organic pollutants in environment of the Pearl river delta,
China: An overview. Chemosphere, 52(9), 1411–1422.
24. Gagliano, E., Sgroi, M., Falciglia, P. P., Vagliasindi, F. G., & Roccaro, P. (2020). Removal of
poly- and perfluoroalkyl substances (PFAS) from water by adsorption: Role of PFAS chain
length, effect of organic matter and challenges in adsorbent regeneration. Water Research,
171, 115381.
25. Saleh, N. B., Khalid, A., Tian, Y., Ayres, C., Sabaraya, I. V., Pietari, J., Hanigan, D., Chowdhury,
I., & Apul, O. G. (2019). Removal of poly- and per-fluoroalkyl substances from aqueous sys-
tems by nano-enabled water treatment strategies. Environmental Science: Water Research &
Technology, 5(2), 198–208.
26. Ebele, A. J., Abdallah, M. A. E., & Harrad, S. (2017). Pharmaceuticals and personal care prod-
ucts (PPCPs) in the freshwater aquatic environment. Emerging Contaminants, 3(1), 1–16.
27. Richardson, B.  J., Lam, P.  K., & Martin, M. (2005). Emerging chemicals of concern:

Pharmaceuticals and personal care products (PPCPs) in Asia, with particular reference to
Southern China. Marine Pollution Bulletin, 50(9), 913–920.
28. Sakkas, V. A., Islam, M. A., Stalikas, C., & Albanis, T. A. (2010). Photocatalytic degradation
using design of experiments: A review and example of the Congo red degradation. Journal of
Hazardous Materials, 175, 33–44.
29. Sin, J. C., Lam, S. M., Mohamed, A. R., & Lee, K. T. (2012). Degrading endocrine disrupt-
ing chemicals from wastewater by TiO2 photocatalysis: A review. International Journal of
Photoenergy, 2012, 185159.
30. Gou, J., Ma, Q., Deng, X., Cui, Y., Zhang, H., Cheng, X., Li, X., Xie, M., & Cheng, Q. (2017).
Fabrication of Ag2O/TiO2–zeolite composite and its enhanced solar light photocatalytic per-
formance and mechanism for degradation of norfloxacin. Chemical Engineering Journal, 308,
818–826.
31. Kurtoglu, M.  E., Longenbach, T., & Gogotsi, Y. (2011). Preventing sodium poisoning of
photocatalytic TiO2 films on glass by metal doping. International Journal of Applied Glass
Science, 2(2), 108–116.
32. Tobajas, M., Belver, C., & Rodriguez, J. J. (2017). Degradation of emerging pollutants in water
under solar irradiation using novel TiO2–ZnO/clay nanoarchitectures. Chemical Engineering
Journal, 309, 596–606.
33. Ahmed, S., Pan, J., Li, D., Tang, P., Shu, X., & Feng, Y. (2020). Growth and removal behavior
of magnesium oxide microspheres towards methyl orange and methylene blue in aqueous solu-
tion. Beilstein Archives, 2020, 20205. https://doi.org/10.3762/bxiv.2020.5.v1.
34. Chaturvedi, G., Kaur, A., Umar, A., Khan, M. A., Algarni, H., & Kansal, S. K. (2020). Removal
of fluoroquinolone drug, levofloxacin, from aqueous phase over iron based MOFs, MIL–100
(Fe). Journal of Solid State Chemistry, 281, 121029.
35. Gupta, G., & Kansal, S. K. (2019). Novel 3-D flower like Bi3O4Cl/BiOCl pn heterojunction
nanocomposite for the degradation of levofloxacin drug in aqueous phase. Process Safety and
Environmental Protection, 128, 342–352.
382 A. Kumar and S. K. Kansal

36. Fung, C.  S., Khan, M., Kumar, A., & Lo, I.  M. C. (2019). Visible–light–driven photocata-
lytic removal of PPCPs using magnetically separable bismuth oxybromo-iodide solid solu-
tions: Mechanisms, pathways, and reusability in real sewage. Separation and Purification
Technology, 216, 102–114.
37. Kaur, N., Shahi, S. K., Shahi, J. S., Sandhu, S., Sharma, R., & Singh, V. (2020). Comprehensive
review and future perspectives of efficient N–doped, Fe–doped and (N, Fe)–codoped titania as
visible light active photocatalysts. Vacuum, 178, 109429.
38. Kumar, A., Khan, M., Zeng, X., & Lo, I. M. (2018). Development of g–C3N4/TiO2/Fe3O4@
SiO2 heterojunction via sol–gel route: A magnetically recyclable direct contact Z–scheme
nanophotocatalyst for enhanced photocatalytic removal of ibuprofen from real sewage effluent
under visible light. Chemical Engineering Journal, 353, 645–656.
39. Eskandarpour, A., Iwai, K., & Asai, S. (2009). Superconducting magnetic filter: Performance,
recovery, and design. IEEE Transactions on Applied Superconductivity, 19(2), 84–95.
40. Gómez-Pastora, J., Dominguez, S., Bringas, E., Rivero, M.  J., Ortiz, I., & Dionysios,

D. D. (2017). Review and perspectives on the use of magnetic nanophotocatalysts (MNPCs) in
water treatment. Chemical Engineering Journal, 310, 407–427.
41. Ahoranta, M., Lehtonen, J., & Mikkonen, R. (2003). Magnet design for superconducting open
gradient magnetic separator. Physica C: Superconductivity, 386, 398–402.
42. Mishra, M., & Chun, D. M. (2015). α–Fe2O3 as a photocatalytic material: A review. Applied
Catalysis A: General, 498, 126–141.
43. Zhang, G. Y., Feng, Y., Xu, Y. Y., Gao, D. Z., & Sun, Y. Q. (2012). Controlled synthesis of
mesoporous α–Fe2O3 nanorods and visible light photocatalytic property. Materials Research
Bulletin, 47(3), 625–630.
44. Asahi, R., Morikawa, T., Irie, H., & Ohwaki, T. (2014). Nitrogen–doped titanium dioxide
as visible–light–sensitive photocatalyst: Designs, developments, and prospects. Chemical
Reviews, 114(19), 9824–9852.
45. Singh, P., Sharma, K., Hasija, V., Sharma, V., Sharma, S., Raizada, P., Singh, M., Saini, A. K.,
Hosseini-Bandegharaei, A., & Thakur, V.  K. (2019). Systematic review on applicability of
magnetic iron oxides–integrated photocatalysts for degradation of organic pollutants in water.
Materials Today Chemistry, 14, 100186.
46. Lamba, R., Umar, A., Mehta, S.  K., & Kansal, S.  K. (2017). Enhanced visible light driven
photocatalytic application of Ag2O decorated ZnO nanorods heterostructures. Separation and
Purification Technology, 183, 341–349.
47. Formal, F. L., Grätzel, M., & Sivula, K. (2010). Controlling photoactivity in ultrathin hematite
films for solar water-splitting. Advanced Functional Materials, 20(7), 1099–1107.
48. Chen, X., Liu, L., Peter, Y. Y., & Mao, S. S. (2011). Increasing solar absorption for photoca-
talysis with black hydrogenated titanium dioxide nanocrystals. Science, 331(6018), 746–750.
49. Kumar, A., Khan, M., He, J., & Lo, I. M. C. (2020b). Visible–light–driven magnetically recy-
clable terephthalic acid functionalized g–C3N4/TiO2 heterojunction nanophotocatalyst for
enhanced degradation of PPCPs. Applied Catalysis B: Environmental, 270, 118898.
50. Álvarez, P.  M., Jaramillo, J., Lopez-Pinero, F., & Plucinski, P.  K. (2010). Preparation and
characterization of magnetic TiO2 nanoparticles and their utilization for the degradation of
emerging pollutants in water. Applied Catalysis B: Environmental, 100(1–2), 338–345.
Chapter 15
Titanate Nanostructures as Potential
Adsorbents for Defluoridation of Water

C. Prathibha, Anjana Biswas, and M. V. Shankar

15.1  Introduction

Water is not only an essential component for life but also a basic building block to
maintain quality of life. Its purity and availability are inextricably linked to global
health and economic development. The presence of several naturally occurring,
anthropogenic, and industry-generated ions such as fluoride, arsenic, nitrate, sul-
fate, iron, manganese, chloride, selenium, heavy metals, and radioactive materials
greatly affects the water quality, leading to health problems. The most significant
inorganic pollutants in groundwater affecting human health at the global scale,
according to the World Health Organization (WHO), are arsenic and fluoride [1].
Fluoride is the only chemical in potable water that can cause varied health effects
depending upon its concentration in dissolved form. It is often described as a
“double-­edged sword” as inadequate ingestion is associated with dental caries,
whereas excessive intake leads to dental, skeletal, and soft tissue fluorosis which has
no cure. A very small amount of fluoride (0.4–1.0 mg/L) is beneficial for bone and
teeth development and dental health. Especially for young children, it promotes
calcification of dental enamel and protects teeth against tooth decay. Therefore, it is
considered as an essential mineral with a narrow margin of safety. Due to these
clinical manifestations caused by drinking fluoride-contaminated water, the WHO
has recommended 1.5 mg/L as the maximum contaminant level (MCL) in drinking
water. Fluorosis due to excessive concentration of fluoride >1.5  mg/L has been
reported in at least 28 countries from South Asia; Africa; the Middle East; North,

C. Prathibha (*) · A. Biswas


Department of Physics, Sri Sathya Sai Institute of Higher Learning, Anantapur Campus,
Anantapur, Andhra Pradesh, India
e-mail: cprathibha@sssihl.edu.in
M. V. Shankar
Nanocatalysis and Solar Fuels Research Laboratory, Department of Materials Science and
Nanotechnology, Yogi Vemana University, Kadapa, Andhra Pradesh, India

© Springer Nature Switzerland AG 2021 383


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_15
384 C. Prathibha et al.

Central, and South America; and Europe as shown in Fig. 15.1 [2]. China and India
are the most affected countries. Defluoridation of drinking water is the only solution
to prevent fluorosis.
Various technologies like chemical precipitation, electrocoagulation, reverse
osmosis, and electrodialysis have been reported to remove fluoride from drinking
water [4, 5]. Despite their unique advantages, these technologies have gained lim-
ited social acceptance due to unaddressed problems such as the high costs, poor
regeneration, interference of other ions, customary replacement of sacrificial elec-
trodes, consumption of electric power, membrane fouling, requirement of experi-
enced operators, and poor water recovery. Scientific evidence recommends
adsorption as the most suitable method as it offers attractive merits such as ease of
operation, simplicity of design, and economics. Adsorption is the adhesion of atoms,
ions, or molecules from liquid to a solid surface as described in Fig. 15.2 [6]. This
process creates a film of the adsorbate on the surface of the adsorbent. The solid
material that provides the surface for adsorption is referred to as adsorbent; the spe-
cies that gets adsorbed is named as adsorbate (fluoride). The success of the adsorp-
tion technique entirely depends on the efficiency of the adsorbents used.
A wide variety of micron-sized adsorbents have been used for the removal of
fluoride from water. These include biosorbents [8], clays [9], soils [10], carbons
[11], zeolites [12], alumina-based materials [13], and synthetic resins [14]. They

Fig. 15.1  Predicted probability of fluoride concentration in the groundwater exceeding the WHO
guideline for drinking water of 1.5  mg  L−1. Reprinted with permission from [3]. Copyright ©
American Chemical Society 2008
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 385

Fig. 15.2  Basic terms of adsorption [7]

have attained limited success due to their poor efficiency and kinetics offered.
Nanomaterials are the best alternatives to the traditional adsorbents. It has been
found that the unique properties of various metal and nonmetal nanomaterials can
be used to develop high capacity and selective sorbents for contaminant removal
from drinking water. This chapter focuses on a conceptual overview of the recent
trends, principles, and applications of advanced nanostructures for the removal of
fluoride from aqueous solution.
Most of the fluoride water filters available in the market for domestic use are very
expensive. A US-based company Crystal Quest manufactures fluoride water filters
for residential use. This system removes ions through a combination of adsorption
and chemical reaction with the media. The price varies from 99 USD to 1566
USD. Aquagear water filtration, USA, has come up with a 1.3-kg water filter pitcher,
which claims to remove fluoride from water. This product is sold in India on differ-
ent online platforms with price per piece starting from 16,000 INR. It is possible to
reduce the cost of these filters with the use of highly efficient cost-effective
nanoadsorbents.

15.2  Nanomaterials as Potential Adsorbents

Nanotechnology has revolutionized the entire scientific and technological fields.


Environmental safety is no exception. One of the most promising and well-­developed
environmental applications of nanotechnology has been in water remediation and
treatment where different nanomaterials can be used to purify water through adsorp-
tion. They have proven themselves to be efficient adsorbents compared to tradition-
ally used materials due to their unique properties at nanoscale. The most vital
properties of these particles which are responsible for their high defluoridation
potential are small size, large surface area, catalytic potential, large number of
active sites, short diffusion route, and high reactivity. These properties fascinated
the scientific community and triggered a great activity in the field of defluoridation.
In the past 10 years, many researchers have devoted their attention to develop low-­
cost and highly efficient nanoadsorbents for the removal of fluoride from aqueous
solution.
386 C. Prathibha et al.

Alumina is the most widely used oxide for defluoridation. Nano-Al2O3 in its
gamma phase (γ-Al2O3) with a very high fluoride adsorption capacity was demon-
strated as the best among all the alumina-based materials [15]. Iron-based oxides
[16] is another well-known adsorbent used for the removal of fluoride from water.
Mixed oxides of calcium, magnesium, and aluminum-based materials in their nano-
form have also been proven to be the highly efficient adsorbents for defluoridation
of water [17, 18].
Adsorption is a surface phenomenon; the surface area of the adsorbent plays a
major role. Large surface area is the property of nanomaterials which can be a use-
ful parameter in enhancing the fluoride adsorption capacity in drinking water treat-
ment. In general, nanomaterials possess higher surface area due to their small size.
It is known that surface properties are determined not only by size but also by shape
[19–21]. For example, sphere and a cube having the same volume, the cube has a
larger surface area than the sphere. Developing nanomaterials in varied morpholo-
gies has been the recent trend in nanoscience research. Moreover, the discovery of
carbon nanotubes stimulated the focus of many researchers toward one-dimensional
nanostructures such as nanorods, nanobelts, and nanotubes due to their exceptional
electrical, mechanical, and chemical properties. Numerous reports [22–24] have
highlighted the use of TiO2-based nanostructures and its composites for sustainable
energy and environmental applications. This chapter describes in detail about TiO2-­
based nanostructures and their application in the field of water treatment for con-
taminant removal.

15.2.1  Titania-Based Nanostructures for Defluoridation

In the past decade, design and fabrication of nanostructures based on metal oxides
have attracted much attention because of their peculiar electronic and optical prop-
erties and their potential applications in the industry and technology [25]. One of
the most intensively studied oxides is titania. Titanium dioxide is proven to be less
toxic [26], and its multifunctional properties enabled them to use in toothpastes; in
food packing; in cosmetic products such as sunscreens, lipsticks, body powders,
soaps, and pearl essence pigments; and also in special pharmaceutics [27]. Various
applications of TiO2 nanoparticles are summarized in Fig. 15.3.
TiO2 exists in three basic phases, namely, anatase, rutile, and brookite [28]. In
general, the anatase phase of TiO2 is preferred due to its higher photocatalytic activ-
ity, photochemically stable nature, nontoxic nature, and more importantly it is rela-
tively inexpensive. The titanates have layered structure which has close structural
resemblance to titanium dioxide, both composed of TiO6 octahedral units connected
by sharing corners and edges as shown in Fig. 15.4.
The three most general approaches to the synthesis of titanate nanostructures are
chemical (template) synthesis, electrochemical approaches (e.g., anodizing of Ti),
and the alkaline hydrothermal method. Alkaline hydrothermal method [30] using
TiO2 (anatase) as a precursor material is one of the reliable ways of producing
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 387

Fig. 15.3  Applications of TiO2 nanoparticles. Reprinted with permission from [27]. Copyright
©Springer Nature Switzerland AG 2019

Fig. 15.4  Crystal structure of the studied TiO2 polymorphs: (a) anatase, (b) rutile, (c) TiO2–B, and
(d) hydrogen titanate (H2Ti3O7). Blue, red, and white spheres represent titanium, oxygen, and
hydrogen atoms, respectively. Reproduced under Creative Commons license, [29]
388 C. Prathibha et al.

Fig. 15.5  Common synthesis routes of TiO2 nanostructures. Reproduced under Creative Commons
license, [32]

Ti-O-based material in varied morphologies such as titanate nanorods (TNRs), tita-


nate nanobelts (TNBs), and titanate nanotubes (TNTs) [28, 31]. Figure 15.5 por-
trays all possible synthesis routes of TiO2 nanostructures.
In the nanostructure form, titania exhibits favorable electronic and optoelectro-
chemical properties with additional features such as high porosity and surface area.
These features greatly enhance the efficiency of the aforementioned applications
and make them suitable for many new applications [33]. The unique structural and
functional properties of TiO2-based nanomaterials have led to breakthroughs in the
field of photocatalysis, photovoltaics, fast-charging lithium-ion batteries, and smart
coatings [28, 34]. Recently, they have also been explored in the field of water treat-
ment for defluoridation of drinking water. Figure  15.6 shows the TEM image of
titanate nanostructures.

15.3  Fluoride Adsorption Studies

Fluoride adsorption studies are generally carried out by potentiometric method by


using the ion-selective electrode. The fluoride adsorption capacity of an adsorbent
is determined by the following expression.

 C − Ce 
qe =  O V
 m 
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 389

Fig. 15.6  TEM images of titanate nanostructures. Reprinted with permission from [30, 35, 36].
Copyright © Springer Nature 2016, © Elsevier 2016, © Taylor and Francis group 2020

where C0 and Ce are the initial and equilibrium fluoride concentration measured in
mg/L. V (L) is the volume of fluoride-containing water treated, and m (g) is the mass
of the adsorbent used. The optimization of the parameters, majorly influencing the
fluoride adsorption process, is evaluated through batch adsorption studies.

15.3.1  E
 ffect of Various Parameters on Fluoride Removal
Efficiency of Titanate Nanostructures

The adsorption of fluoride ions by an adsorbent depends upon different aspects, for
example, surface area, dose, and the isoelectric point (IEP) of an adsorbent. The
initial concentration of fluoride ion, contact time, and pH of the solution equally
influence the adsorption efficiency of an adsorbent. The optimization of these
parameters plays an important role to identify the best operating parameters that
would give the maximum adsorption capacity of the adsorbent. This becomes spe-
cifically essential while estimating the applicability of the adsorbent and designing
a product for real-time application such as a filter for remediating groundwater and
390 C. Prathibha et al.

wastewater. Further, knowing the best operating parameters, one could fine-tune the
operating conditions in order to enhance the adsorption capacity. The optimization
is usually done using the batch adsorption experiments. An ideal adsorbent would
have a wide pH window for adsorption and high fluoride uptake in a short contact
time with low adsorbent dosage.

Surface Area

Higher surface area offers a major portion of its particles to be present at the surface
which in turn increase the number of active sites for fluoride adsorption [37]. As
shown in Fig. 15.7, the particles forming the bulk of any material experiences equal
forces from all sides and, hence they do not have any unbalanced force. On the other
hand, the particles on the surface of the materials experience unbalanced forces,
which form the driving force for attracting the adsorbate ions/molecules. As the
surface area of the material increases, the driving force for adsorption also increases.
Surface area of an adsorbent greatly influences its adsorption efficiency, since
adsorption is a surface phenomenon.
Among the three variants of titanate nanostructures, TNTs possess the highest
surface area which could be attributed to their hollow tubular nature. Comparing the
adsorption capacities of the three one-dimensional nanostructures, titanate nano-
tubes (TNTs), titanate nanobelts (TNBs), and titanate nanorods (TNRs), it can be
observed that fluoride adsorption capacity follows the trend TNT > TNB > TNR. The
higher adsorption capacity of TNT compared to the TNB and TNR is mainly due to
higher surface area as a result of its tubular morphology. The surface area of TNRs,
TNBs, and TNTs is reported to be 29, 38, and 282  m2/g, respectively, which is
nearly 5, 6, and 60 times higher than the precursor material, TiO2 fine particles
(5 m2/g) [30, 35]. TNT with higher surface area offers a greater number of adsorp-
tion sites on its surface and hence the higher adsorption capacity. Further, the TNR
with the lowest surface area has the least number of adsorption sites available on the

Fig. 15.7  Driving force for adsorption: unbalanced forces of surface atoms
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 391

surface and therefore the least adsorption capacity. However, all the three one-­
dimensional morphologies of titanates have superior fluoride adsorption compared
to TiO2 with spherical morphology. It could be seen that there exists a perfect cor-
relation between fluoride adsorption efficiency and the specific surface area of the
adsorbent. Titanate nanostructures demonstrate the effect of morphology on fluo-
ride adsorption capacity. Hence, for effective adsorption to take place, adsorbents
with higher surface areas are required.

Dose of an Adsorbent Used

Dosage is the quantity of adsorbent in g/L that is added to the fluoride-containing


water. An ideal adsorbent should exhibit maximum adsorption capacity with mini-
mum dosage. In general, adsorption capacity can be increased with increase in
adsorbent dose. This is due to the increased number of available adsorption sites
which results in enhanced adsorption capacity. This trend in variation of adsorption
capacity with adsorbent dosage was reported by various researchers [30, 35, 38].
However, to be a cost-effective adsorbent, the optimum dose for defluoridation
ought to be as low as possible. This requirement was satisfied by the titanate nano-
structures. TNT is reported to have reduced higher fluoride concentration to WHO
limits with the use of as low as 1 g/L dosage. This was possible due to the high
surface area of TNTs by virtue of its tubular morphology [35]. In general, nanoma-
terials with very low adsorbent dosage have proven to exhibit much higher fluoride
adsorption capacity when compared to bulk materials. Figure 15.8 shows the com-
parison. Rajan et al. [39] reported the usage of 15 g/L of zirconium-impregnated
walnut shell carbon to treat water with 3 mg/L of fluoride which is a higher dosage
when compared to dosage of titanates used for defluoridation.

Fig. 15.8  Variation of adsorption capacity with adsorbent dosage using (a) natural clay (pH 5.8,
initial fluoride concentration 5 mg/L, contact time 3 h) and (b) titanate nanotubes (pH 2, initial
fluoride concentration 10  mg/L, contact time 2  h). Reprinted with permission from [9, 35].
Copyright © Académie des sciences 2018, © Elsevier 2016
392 C. Prathibha et al.

Contact Time

Contact time is another vital factor in fluoride adsorption process by an adsorbent.


Initially, the adsorption is rapid due to availability of vacant active sites on the sur-
face of an adsorbent; as the equilibrium approaches, the number of free sites on the
adsorbent surface decreases, and the adsorption capacity attains saturation.
Equilibrium time is the minimum time for which adsorbent needs to be in contact
with fluoride ions in order to obtain maximum fluoride removal. Numerous reports
[35, 38, 40] indicate the quick fluoride adsorption process, and lesser equilibrium
time  taken by one-dimensional nanostructures compared to conventionally used
micron-sized adsorbents. This is illustrated in Fig. 15.9. This could be attributed to
the fact that nanomaterials have much greater number of adsorbent sites readily
available for the uptake of adsorbate ions compared to their bulk counterparts. This
is due to their high surface area which in turn ensures quick adsorption. Hence,
researchers are working toward developing adsorbents which exhibit instantaneous
adsorption using nanomaterials, in which equilibrium adsorption occurs within a
few minutes.

Isoelectric Point and pH of the Solution

Zeta potential is a measure of effective surface charge of nanoadsorbent which var-


ies with the pH of adsorbent solution. The isoelectric point (IEP) is the pH at which
the zeta potential of the adsorbent is zero. At IEP, both positive and negative charges
are well compensated that leads to no excess surface charges on the surface of tita-
nate nanostructures.
At this pH the number of positive surface sites equals the number of negative
surface sites. In addition, at all pH < IEP, the surface charge of an adsorbent is posi-
tive, whereas at all pH > IEP the surface charge of the adsorbent is negative. This is

Fig. 15.9  Comparison of equilibrium contact time using (a) nanomaterial (CeO2–ZrO2 nano-
cages) (adsorbent dosage 0.2 g/L, pH = 4) and (b) bulk material (zirconium I-impregnated coconut
shell fiber) (adsorbent dosage 10 g/L, pH = 4). Reprinted with permission from [41, 42]. Copyright
© Elsevier 2013, © Taylor and Francis Group LLC 2008
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 393

Table 15.1  Fluoride adsorption efficiencies of Ti-O-based adsorbents and pH conditions at which
they exhibit maximum efficiency
pH at which adsorbent Maximum
showed maximum adsorption adsorption capacity
Adsorbent capacity (mg/g) References
TiO2 1 100.7 [47]
TiO2-SiO2 nanocomposite 1 152.2 [47]
Titanate nanotubes 2–3 58 [35]
Anatase TiO2 3–4 32.15 [48]
TiO2 2–4 0.85 [49]
Microbially synthesized
Aluminum titanate 3–9 0.85 [50]
Magnesium titanate 3–11 0.029 [51]
FZTNT: Fe(III) and Zr(IV) 7 229 [45]
surface-functionalized TNT

an important parameter which determines the pH sensitivity of an adsorbent in


adsorption process. Fluoride adsorption is favorable when pH of the solution is less
than IEP of an adsorbent as it favors electrostatic attraction between the positively
charged surface and negative fluoride ion, whereas at pH greater than IEP the reverse
process occurs, as the negative adsorbent surface now repels the negatively charged
fluoride ions. Most of the titanates reported in literature exhibit favorable fluoride
adsorption in acidic conditions due to their lower IEP values (Table 15.1).

 luoride Adsorption Mechanism by One-Dimensional


F
Titanate Nanostructures

The protonic trititanate nanostructures are made up of Ti and O, stacked up into


multiple layers in its crystal structure with ion-exchangeable H+ and OH−. The
mechanism of fluoride adsorption by titanate nanostructures is well understood
based on the abundant availability of Ti-OH ions on the titanate nanostructures’
surface. These hydroxyl ions act as the active sites for effective adsorption of fluo-
ride ions from water. The similar ionic radii of F− and OH− ions indicate the possi-
bility of fluoride ions from the aqueous media undergoing ion exchange with the
surface OH− ions [30]. “Ion exchange is a stoichiometric process where any counter
ion leaving the ion exchanger surface is replaced by an equivalent number of moles
of another counter ion to maintain electro-neutrality of the ion exchanger” [43].
Figure 15.10 depicts the ion-exchange mechanism by titanate nanostructures.
The mechanism is usually supported by experimental evidences such as Fourier
transform infrared spectroscopy (FTIR) and x-ray photoelectron spectroscopy
(XPS) of the adsorbent before and after fluoride adsorption. Using both the charac-
terization techniques, the presence of M-OH (M = metal) peak could be observed
which confirms the presence of -OH on adsorbent surface. The decrease in intensity
of the M-OH peak in FTIR spectra of post-fluoride adsorption could be ascribed to
394 C. Prathibha et al.

Fig. 15.10  Schematic representation of ion-exchange mechanism 1-D titanate nanostructures and
spherical nanoparticle. Reprinted with permission from [35]. Copyright ©Elsevier 2016

the successful ion exchange between the hydroxyl group and the fluoride ions [35,
44–46]. In addition, the shift in the high-resolution peaks of Ti in the XPS spectra
of the titanates toward higher binding energy in post-fluoride adsorption proves the
successful uptake of fluoride ions by the adsorbent [44, 45].
The initial pH of fluoride-containing solution also greatly affects the surface
charge of adsorbent, which is demonstrated using isoelectric point (IEP). The con-
centrations of protons or surface hydroxyl groups which are responsible for fluoride
uptake process by an adsorbent depend on its IEP and pH of the fluoride-containing
solution. Therefore, adsorption mechanism is well understood based on the follow-
ing two possible reactions at the oxide water interface:

Ti − OH + H 2 O → Ti − OH 2+ + OH − pH < IEP (15.1)

Ti − OH + H 2 O → Ti − O − + H 3 O + pH > IEP (15.2)

Equilibrium (15.1) is favorable when pH of the solution is lesser than IEP,


whereas equilibrium (15.2) is more favorable when pH of the solution is greater
than IEP. However, in any condition whether acidic or basic, Ti − OH 2+, Ti − OH,
and Ti − O− are present but in different concentrations which determine the adsorp-
tion capacity. In acidic media (pH < IEP), more of Ti − OH 2+-positive ions are avail-
able on adsorbent surface; hence it favors the uptake of negatively charged fluoride
ions from the water as per the equation given below [30, 35]:

Ti − OH 2+ + F − → Ti − F + H 2 O (15.3)
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 395

Therefore, the fluoride adsorption by the titanate nanostructures is majorly due


to ion-exchange mechanism and electrostatic attraction.

15.4  Surface Functionalization

Though the adsorbent, titanate nanotube (TNT), exhibits a very high fluoride
adsorption capacity, it is highly sensitive to pH of water, and its efficiency is limited
to acidic medium at pH 2. Hence it could not be used for removing fluoride from
groundwater samples at neutral pH. This was due to the negative surface charge of
TNT which repelled the negative fluoride ion in alkaline conditions, and the same
was confirmed by zeta a potential measurement. Surface functionalization with sev-
eral cations such as Zr4+, Fe3+, La3+, Ca2+, Cu2+, Ag+, Ni2+, and Co2+ in varied concen-
trations and combinations is one of the ways to address this problem [43]. Surface
modification makes the TNT surface more positive or less negative at neutral pH so
that the material offers its efficiency for fluoride adsorption at neutral pH as well.
Recently a shift in IEP of TNT as a result of surface functionalization with Fe and
Zr (FZTNT) is reported [45].
Figure 15.11, the plot of zeta potential vs pH of TNT and FZTNT, clearly indi-
cates a shift in IEP with varied molar ratios of Fe and Zr. Figure 15.12 gives the fluo-
ride adsorption capacity of TNT and FZTNT (with varied molar ratios of Fe and Zr)
as a function of pH.

Fig. 15.11  The plot of zeta potential vs pH of TNT and FZTNT (varied molar ratios of Fe and Zr).
Reprinted with permission from [45]. Copyright © Elsevier 2020
396 C. Prathibha et al.

Fig. 15.12  Fluoride ion adsorption of TNT and FZTNT as a function of initial pH (varied molar
ratios of Fe and Zr). Reprinted with permission from [45]. Copyright © Elsevier 2020

The fluoride adsorption capacity of FZTNT has a large pH window compared to


TNT. In addition, it has a considerably higher adsorption at neutral pH, when com-
pared to TNT, confirming the results obtained from the zeta potential measurements.
Therefore, with a suitable surface functionalization, titanate nanostructure could be
used to treat fluoride-contaminated water at neutral pH. This enables them to be the
potential fluoride adsorbents to treat groundwater contaminated with excessive fluo-
ride. Moreover, surface functionalization increases the number of OH− ions on the
surface of FZTNT which enhances ion-exchange mechanism. As a result, FZTNT
has four times higher fluoride adsorption capability than TNT. Table 15.1 gives the
comparative study of fluoride adsorption capacities of Ti-O-based materials.

15.5  Conclusion and Scope for Future Research

In summary, producing nanomaterials in varied morphologies is a novel approach to


develop nanoadsorbents for defluoridation of water. This chapter reported morpho-
logical effects of titanate nanostructures on efficient defluoridation. Hydrothermally
synthesized titanate nanostructures exhibit enhanced fluoride adsorption capacity
compared to the traditionally used micro-/nano-size adsorbents. The high adsorp-
tion capacity of these nanostructures is due to the availability of abundant active
sites as a result of their morphology and high surface area. Ion-exchange mecha-
nism and electrostatic attraction at lower pH explain the interaction between
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 397

fluoride ion and titanate nanostructure during adsorption process. It is possible to


remove the fluoride from water with the use of lesser amounts of these nanostruc-
tures in a short duration of time. However, despite all their merits, they could not be
used for treating groundwater at neutral pH due to their limited efficiency in acidic
medium. This chapter reports an efficient way to transform titanate nanostructures
so that they can offer their efficiency at neutral pH to treat groundwater as well.
Surface modification with Fe(III) and Zr(IV) ions transforms TNT to a remarkable
defluoridating agent which can be used to treat groundwater at neutral pH. Surface
functionalization retains all the merits of titanate nanostructures while transforming
their surface suitable for defluoridation at neutral pH.
Development of inexpensive methods to produce titanate nanostructures is one
of the challenges of future research for their use in large-scale defluoridation pro-
cess. In addition, there is a need for an insight study of the reusability and sturdiness
of surface functionalized nanostructures overtime in aqueous media.

References

1. Al-Sulaiman, S. A. A.-W. (2012). Chemical safety of drinking-water: Assessing priorities for
risk management. International Journal of Environmental Studies, 69, 1001–1001. https://doi.
org/10.1080/00207233.2011.565947.
2. Fawell, J., Bailey, K., Chilton, J., Dahi, E., Fewtrell, L., & Magara, Y. (2006). WHO Drinking-­
Water Quality Series, Fluoride in Drinking Water (2006.).  IWA Publishing, U.K. https://www.
iwapublishing.com/books/9781900222969/fluoride-drinking-water
3. Amini, M., Mueller, K., Abbaspour, K. C., Rosenberg, T., Afyuni, M., Møller, K. N., Sarr, M.,
& Johnson, C. A. (2008). Statistical modeling of global geogenic fluoride contamination in
groundwaters. Environmental Science & Technology, 42, 3662–3668. https://doi.org/10.1021/
es071958y.
4. Habuda-Stanić, M., Ravančić, M., & Flanagan, A. (2014). A review on adsorption of fluoride
from aqueous solution. Materials (Basel), 7, 6317–6366. https://doi.org/10.3390/ma7096317.
5. Akafu, T., Chimdi, A., & Gomoro, K. (2019). Removal of fluoride from drinking water by
sorption using diatomite modified with aluminum hydroxide. Journal of Analytical Methods
in Chemistry, 2019, 4831926. https://doi.org/10.1155/2019/4831926.
6. Singh, K., Lataye, D. H., & Wasewar, K. L. (2016). Removal of fluoride from aqueous solu-
tion by using low-cost sugarcane bagasse: Kinetic study and equilibrium isotherm analyses.
Journal of Hazardous, Toxic, and Radioactive Waste, 20, 04015024. https://doi.org/10.1061/
(ASCE)HZ.2153-­5515.0000309.
7. Worch, E. (2012). Adsorption technology in water treatment. https://doi.
org/10.1515/9783110240238.
8. Amin, F., Talpur, F.  N., Balouch, A., Surhio, M.  A., & Bhutto, M.  A. (2015). Biosorption
of fluoride from aqueous solution by white—Rot fungus Pleurotus eryngii ATCC 90888.
Environmental Nanotechnology, Monitoring and Management, 3, 30–37. https://doi.
org/10.1016/j.enmm.2014.11.003.
9. Nabbou, N., Belhachemi, M., Boumelik, M., Merzougui, T., Lahcene, D., Harek, Y., Zorpas,
A.  A., & Jeguirim, M. (2019). Removal of fluoride from groundwater using natural clay
(kaolinite): Optimization of adsorption conditions. Comptes Rendus Chimie, 22, 105–112.
https://doi.org/10.1016/j.crci.2018.09.010.
398 C. Prathibha et al.

10. Iriel, A., Bruneel, S. P., Schenone, N., & Cirelli, A. F. (2018). The removal of fluoride from
aqueous solution by a lateritic soil adsorption: Kinetic and equilibrium studies. Ecotoxicology
and Environmental Safety, 149, 166–172. https://doi.org/10.1016/j.ecoenv.2017.11.016.
11. Araga, R., Soni, S., & Sharma, C.  S. (2017). Fluoride adsorption from aqueous solu-

tion using activated carbon obtained from KOH-treated jamun (Syzygium cumini) seed.
Journal of Environmental Chemical Engineering, 5, 5608–5616. https://doi.org/10.1016/j.
jece.2017.10.023.
12. Sun, Y., Fang, Q., Dong, J., Cheng, X., & Xu, J. (2011). Removal of fluoride from drinking
water by natural stilbite zeolite modified with Fe(III). Desalination, 277, 121–127. https://doi.
org/10.1016/j.desal.2011.04.013.
13. Sujana, M. G., Thakur, R. S., & Rao, S. B. (1998). Removal of fluoride from aqueous solu-
tion by using alum sludge. Journal of Colloid and Interface Science, 206, 94–101. https://doi.
org/10.1006/jcis.1998.5611.
14. Paudyal, H., Inoue, K., Kawakita, H., Ohto, K., Kamata, H., & Alam, S. (2018). Removal
of fluoride by effectively using spent cation exchange resin. Journal of Material Cycles and
Waste Management, 20, 975–984. https://doi.org/10.1007/s10163-­017-­0659-­4.
15. Chinnakoti, P., Chunduri, A.  L. A., Vankayala, R.  K., Patnaik, S., & Kamisetti, V. (2017).
Enhanced fluoride adsorption by nano crystalline γ-alumina: Adsorption kinetics, isotherm
modeling and thermodynamic studies. Applied Water Science, 7, 2413–2423. https://doi.
org/10.1007/s13201-­016-­0437-­9.
16. Prathna, T. C., Sharma, S. K., & Kennedy, M. (2017). Development of iron oxide nanoparticle
adsorbents for arsenic and fluoride removal. Desalination and Water Treatment, 67, 187–195.
https://doi.org/10.5004/dwt.2017.20464.
17. Zhang, C., Li, Y., Jiang, Y., & Wang, T. J. (2017). Size-dependent fluoride removal performance
of a magnetic Fe3O4@Fe-Ti adsorbent and its defluoridation in a fluidized bed. Industrial and
Engineering Chemistry Research, 56, 2425–2432. https://doi.org/10.1021/acs.iecr.6b03856.
18. Liu, L., Cui, Z., Ma, Q., Cui, W., & Zhang, X. (2016). One-step synthesis of magnetic iron-­
aluminum oxide/graphene oxide nanoparticles as a selective adsorbent for fluoride removal
from aqueous solution. RSC Advances, 6, 10783–10791. https://doi.org/10.1039/c5ra23676b.
19. Wang, Z., Lv, K., Wang, G., Deng, K., & Tang, D. (2010). Study on the shape control and pho-
tocatalytic activity of high-energy anatase titania. Applied Catalysis B: Environmental, 100,
378–385. https://doi.org/10.1016/j.apcatb.2010.08.014.
20. Mino, L., Spoto, G., & Ferrari, A. M. (2014). CO2 capture by TiO2 anatase surfaces: A com-
bined DFT and FTIR study. Journal of Physical Chemistry C, 118, 25016–25026. https://doi.
org/10.1021/jp507443k.
21. Mino, L., Pellegrino, F., Rades, S., Radnik, J., Hodoroaba, V.  D., Spoto, G., Maurino, V.,
& Martra, G. (2018). Beyond shape engineering of TiO2 nanoparticles: Post-synthesis treat-
ment dependence of surface hydration, hydroxylation, Lewis acidity and photocatalytic activ-
ity of TiO2 anatase nanoparticles with dominant {001} or {101} facets. ACS Applied Nano
Materials, 1, 5355–5365. https://doi.org/10.1021/acsanm.8b01477.
22. Wang, S., Quan, W., Zhu, Z., Yang, Y., Liu, Q., Ren, Y., Zhang, X., Xu, R., Hong, Y., Zhang, Z.,
Amine, K., Tang, Z., Lu, J., & Li, J. (2017). Lithium titanate hydrates with superfast and sta-
ble cycling in lithium ion batteries. Nature Communications, 8, 627. https://doi.org/10.1038/
s41467-­017-­00574-­9.
23. Henao, J., Pacheco, Y., Sotelo, O., Casales, M., & Martinez-Gómez, L. (2019). Lanthanum
titanate nanometric powder potentially for rechargeable Ni-batteries: Synthesis and elec-
trochemical hydrogen storage. Journal of Materials Research and Technology, 8, 759–765.
https://doi.org/10.1016/j.jmrt.2018.05.019.
24. Fleischmann, S., Pfeifer, K., Widmaier, M., Shim, H., Budak, Ö., & Presser, V. (2019).
Understanding interlayer deprotonation of hydrogen titanium oxide for high-power elec-
trochemical energy storage. ACS Applied Energy Materials, 2, 3633–3641. https://doi.
org/10.1021/acsaem.9b00363.
15  Titanate Nanostructures as Potential Adsorbents for Defluoridation of Water 399

25. Pan, Z. W., Dai, Z. R., & Wang, Z. L. (2001). Nanobelts of semiconducting oxides. Science,
291, 1947–1949. https://doi.org/10.1126/science.1058120.
26. Maiti, A., Mishra, S., & Chaudhary, M. (2018). Nanoscale materials for arsenic removal from
water. Elsevier Inc.. https://doi.org/10.1016/B978-­0-­12-­813926-­4.00032-­X.
27. Waghmode, M. S., Gunjal, A. B., Mulla, J. A., Patil, N. N., & Nawani, N. N. (2019). Studies
on the titanium dioxide nanoparticles: Biosynthesis, applications and remediation. SN Applied
Sciences, 1, 1–9. https://doi.org/10.1007/s42452-­019-­0337-­3.
28. Ali, I., Suhail, M., Alothman, Z. A., & Alwarthan, A. (2018). Recent advances in syntheses,
properties and applications of TiO2 nanostructures. RSC Advances, 8, 30125–30147. https://
doi.org/10.1039/c8ra06517a.
29. German, E., Faccio, R., & Mombrú, A. W. (2017). A DFT + U study on structural, electronic,
vibrational and thermodynamic properties of TiO2 polymorphs and hydrogen titanate: Tuning
the hubbard ‘U-term’. J. Phys. Commun., 1, 055006. https://doi.org/10.1088/2399-­6528/aa8573.
30. Chinnakoti, P., Kurdekar, A. D., Avinash Chundur, L. A., Aditha, S., Biswas, A., Mthukonda,
S. V., & Kamisetti, V. (2020). Titanate nanobelts—A promising nanosorbent for defluoridation
of drinking water. Sep. Sci. Technol., 55, 1023–1035. https://doi.org/10.1080/01496395.201
9.1580733.
31. Kim, G. S., Kim, Y. S., Seo, H. K., & Shin, H. S. (2006). Hydrothermal synthesis of titanate
nanotubes followed by electrodeposition process. Korean Journal of Chemical Engineering,
23, 1037–1045. https://doi.org/10.1007/s11814-­006-­0027-­x.
32. Shaheed, M. A., & Hussein, F. H. (2014). Preparation and applications of titanium dioxide and
zinc oxide nanoparticles. Journal of Environmental Analytical Chemistry, 02, 9–11. https://doi.
org/10.4172/jreac.1000e109.
33. Lee, K., Mazare, A., & Schmuki, P. (2014). One-dimensional titanium dioxide nanomaterials:
Nanotubes. Chemical Reviews, 114, 9385–9454. https://doi.org/10.1021/cr500061m.
34. Wang, X., Li, Z., Shi, J., & Yu, Y. (2014). One-dimensional titanium dioxide nanomateri-
als: Nanowires, nanorods, and nanobelts. Chemical Reviews, 114, 9346–9384. https://doi.
org/10.1021/cr400633s.
35. Chinnakoti, P., Vankayala, R. K., Chunduri, A. L. A., Nagappagari, L. R., Muthukonda, S. V.,
& Kamisetti, V. (2016). Trititanate nanotubes as highly efficient adsorbent for fluoride removal
from water: Adsorption performance and uptake mechanism. Journal of Environmental
Chemical Engineering, 4, 4754–4768. https://doi.org/10.1016/j.jece.2016.11.007.
36. Lakshmana Reddy, N. L., Praveen Kumar, D. P., & Shankar, M. V. (2016). Co-catalyst free
Titanate Nanorods for improved Hydrogen production under solar light irradiation. Journal of
Chemical Sciences, 128, 649–656. https://doi.org/10.1007/s12039-­016-­1061-­9.
37. Vickers, N. J. (2017). Animal communication: When I’m calling you, will you answer too?
Current Biology, 27, R713–R715. https://doi.org/10.1016/j.cub.2017.05.064.
38. Zhang, Y., Qian, Y., Li, W., Gao, X., & Pan, B. (2019). Fluoride uptake by three lanthanum
based nanomaterials: Behavior and mechanism dependent upon lanthanum species. Science of
the Total Environment, 683, 609–616. https://doi.org/10.1016/j.scitotenv.2019.05.185.
39. Rajan, M., & Alagumuthu, G. (2013). Study of fluoride affinity by zirconium impregnated
walnut shell carbon in aqueous phase: Kinetic and isotherm evaluation. Journal of Chemistry,
2013, 235048. https://doi.org/10.1155/2013/235048.
40. Fernando, M.  S., Wimalasiri, A.  K. D.  V. K., Ratnayake, S.  P., Jayasinghe, J.  M. A.  R. B.,
William, G. R., Dissanayake, D. P., De Silva, K. M. N., & De Silva, R. M. (2019). Improved
nanocomposite of montmorillonite and hydroxyapatite for defluoridation of water. RSC
Advances, 9, 35588–35598. https://doi.org/10.1039/c9ra03981c.
41. Wang, J., Xu, W., Chen, L., Jia, Y., Wang, L., Huang, X. J., & Liu, J. (2013). Excellent fluoride
removal performance by CeO2-ZrO2 nanocages in water environment. Chemical Engineering
Journal, 231, 198–205. https://doi.org/10.1016/j.cej.2013.07.022.
42. Sathish, R.  S., Raju, N.  S. R., Raju, G.  S., Rao, G.  N., Kumar, K.  A., & Janardhana,
C. (2007). Equilibrium and kinetic studies for fluoride adsorption from water on zirconium
400 C. Prathibha et al.

i­mpregnated coconut shell carbon. Separation Science and Technology, 42, 769–788. https://
doi.org/10.1080/01496390601070067.
43. Loganathan, P., Vigneswaran, S., Kandasamy, J., & Naidu, R. (2013). Defluoridation of drink-
ing water using adsorption processes. Journal of Hazardous Materials, 248–249, 1–19. https://
doi.org/10.1016/j.jhazmat.2012.12.043.
44. Lin, J., Wu, Y., Khayambashi, A., Wang, X., & Wei, Y. (2018). Preparation of a novel
CeO2/SiO2 adsorbent and its adsorption behavior for fluoride ion. Adsorption Science and
Technology, 36, 743–761. https://doi.org/10.1177/0263617417721588.
45. Biswas, A., & Prathibha, C. (2020). Fe(III) and Zr(IV) surface functionalized 1-D hydrogen
titanate nanotubes for remediating fluoride from water at neutral pH. Journal of Water Process
Engineering, 37, 101331. https://doi.org/10.1016/j.jwpe.2020.101331.
46. Dou, X., Mohan, D., Pittman, C.  U., & Yang, S. (2012). Remediating fluoride from water
using hydrous zirconium oxide. Chemical Engineering Journal, 198–199, 236–245. https://
doi.org/10.1016/j.cej.2012.05.084.
47. Zeng, Y., Xue, Y., Liang, S., & Zhang, J. (2017). Removal of fluoride from aqueous solution
by TiO2 and TiO2–SiO2 nanocomposite. Chemical Speciation and Bioavailability, 29, 25–32.
https://doi.org/10.1080/09542299.2016.1269617.
48. Deng, H., Zhang, K., & Wang, X. (2016). Synthesis of titanate nanoparticles in low tempera-
ture hydrolysis and adsorption of arsenate (V) and fluoride. Desalination and Water Treatment,
57, 9409–9419. https://doi.org/10.1080/19443994.2015.1030707.
49. Suriyaraj, S.  P., Vijayaraghavan, T., Biji, P., & Selvakumar, R. (2014). Adsorption of fluo-
ride from aqueous solution using different phases of microbially synthesized TiO2 nanopar-
ticles. Journal of Environmental Chemical Engineering, 2, 444–454. https://doi.org/10.1016/j.
jece.2014.01.013.
50. Karthikeyan, M., & Elango, K. P. (2009). Removal of fluoride from water using aluminium
containing compounds. Journal of Environmental Sciences, 21, 1513–1518. https://doi.
org/10.1016/S1001-­0742(08)62448-­1.
51. Gopal, V., & Elango, K. P. (2010). Studies on defluoridation of water using magnesium tita-
nate. Indian Journal of Chemical Technology, 17, 28–33.
Chapter 16
Photocatalytic Water Pollutant Treatment:
Fundamental, Analysis and Benchmarking

Katherine Rebecca Davies, Ben Jones, Chiaki Terashima, Akira Fujishima,


and Sudhagar Pitchaimuthu

16.1  Introduction

The population of the world is constantly growing with the expectation that by 2050
there will be 9.5 billion people on the planet. This increases the demand for food,
clothes, power and drinking water; thus there is a huge pressure on the water treat-
ment industry to produce a significant amount of water to sustain us. This is a mas-
sive concern as currently there are 1.2 billion people without safe drinking water
and 2.6 billion with little or no sanitation; thus millions of people die annually
through contamination in water. Therefore, developing affordable water cleaning/
treatment systems at a low cost are is highly critical for developing countries.
Currently, there are many water cleaning technologies used to help clean freshwater
sources for use in the industry. These technologies include coagulation and floccula-
tion, adsorption, ion-exchange, membrane filtration, electrodialysis, biological
treatment and ozonation. All these technologies have demonstrated effective remov-
als from wastewater with ion exchange and reserve osmosis being the most popular
choice, due to their greater capacity and more effective removals, while for dye
adsorption it is very popular due to the high decolourization efficiency, high affinity
and capability and low-cost adsorbent material. However, these technologies also
have disadvantages, namely, high costs, high waste production (from membrane
fouling), pollutants not 100% removed, high energy demand and some are complex,
thus requiring trained operators.

K. R. Davies · B. Jones · S. Pitchaimuthu (*)


Multi-functional Photocatalyst and Coatings Group, SPECIFIC, College of Engineering,
Swansea University (Bay Campus), Swansea, Wales, UK
e-mail: S.Pitchaimuthu@swansea.ac.uk
C. Terashima · A. Fujishima
Photocatalysis International Research Center, Tokyo University of Science,
Nodashi, Chiba ken, Japan

© Springer Nature Switzerland AG 2021 401


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_16
402 K. R. Davies et al.

Researchers have now been looking at more sustainable processes which can
achieve effective pollutant removal with less of an environmental impact. One of
these processes is photocatalysis which has proven a popular choice since 1972, and
there are currently more than 62,000 publications (Fig. 16.1a; source, Scopus search
engine; keyword, photocatalysis) in this area. In this context, titanium dioxide
(TiO2) is a promising choice (more than 28,000 publications) for photocatalysis due
to its nontoxic, highly photoreactive (in UV light), low-cost, photostable, chemi-
cally and biologically inert [1], wide bandgap and dielectric properties. Though
photocatalysis technique has resulted in high removals of organic pollutants (dyes,
pesticides and pharmaceutical industry waste), it does have challenges such as (a)
mass production of TiO2 and (b) other photocatalyst materials consisting of toxic
components, thus having an environmental impact if they leach from the process.
Furthermore, it can increase the operating cost due the separation of the photocata-
lyst particles after the water treatment process. However, immobilization on station-
ary electrodes can help avoid this step and help integrate this process into the
industry.
Apart from the technical challenges, photocatalysis has been gathering great
interest in recent years. A significant number of review articles and books (refer
Fig. 16.1b) have been published on photocatalysis topics. However, the fundamen-
tals on understanding photocatalysis technique in pollutant environment, analytical
techniques and their operating principles and new forms of value-added recovery
materials from photocatalysis process need to be discussed in detail. Therefore, this
chapter will focus on utilizing semiconductor photocatalyst to remove a wide range
of organic pollutants in water and how to analyse these target pollutants quantita-
tively after the treatment process. Importantly, three main pollutants which can be
seen in freshwater environment have been considered, namely, dyes, pesticides and
pharmaceutical industry waste. For easy understanding we have summarized recent
advances in TiO2-based photocatalysis as is well documented.

Fig. 16.1  Publication record on “photocatalysis” topic (a) year wise and (b) documents by type.
Note that this data was collected from scopus.com on February 2020, using keyword
“photocatalysis”
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 403

16.2  Theory

16.2.1  Fundamentals in Photocatalysis

Since the ground-breaking discovery of electrochemical photolysis of water by


Fujishima and Honda [2] in year 1972, photocatalysis research around the world has
accelerated significantly. This work explains the interaction of light and semicon-
ductor in oxidation of water molecules, which is analogy to nature photosynthesis
process opens a new type of semiconductors and nanoscale architectures in photo-
catalysis research. The photocatalyst is placed in the solution and is irradiated by a
light (with energy more than the bandgap of the semiconductor), thus producing an
electron-hole pair due to the electron being excited to the conduction band from the
valance band [3]. The resultant photocharge carriers migrate to the surface of the
semiconductor which causes redox reactions to occur at the interface. The organic
pollutant can be directly oxidized by holes in valance band, the hydroxyl radicals
(produced from water-splitting reactions) or the superoxide (O2·) radicals. The
superoxide radicals are producing by the reduction of dissolved oxygen by elec-
trons. Recombination of the photocharge carriers, i.e. electrons and holes, can occur
in the trapping sites due to the photogenerated carriers both being present on the
surface of the photocatalyst [4].
The performance of photocatalysis is dependent on the adsorption of the pollut-
ant onto the surface of the semiconductor [5]. The more pollutant absorbed on the
surface, the higher the amount of pollutants degraded. Therefore, it has been dem-
onstrated that nanosized TiO2 (anatase form) provides better results than bulk TiO2,
due to the bigger surface to volume in a nanoscale [6]. The schematic illustration of
photocatalysis mechanism in organic dye pollutant treatment is presented in
Fig. 16.2a, b.
Photocatalysis is a prematured technology, and the mechanism has been fully
developed for the degradation of pollutants and reaction paths below [8, 9]:

(
TiO2 + hν ( UV ) → TiO2 e CB− + h VB+ ) (16.1)

TiO 2 ( h VB + ) + H 2 O → TiO 2 + H + + OH . (16.2)


TiO 2 ( h VB + ) + OH − → TiO 2 + OH . (16.3)


TiO 2 ( eCB − ) + O 2 → TiO 2 + O 2 − . (16.4)



O 2 − . + H + → HO 2 . (16.5)

Dye + OH . → degradation products (16.6)

Dye + h VB+ → oxidation products (16.7)


404 K. R. Davies et al.

Fig. 16.2  Schematic illustration of (a) photocatalytic dye pollutant degradation under light illumi-
nation, (b) charge recombination routes and (c) energetic structure of semiconductors suitable for
water oxidation/reduction and OH•/O2−• production (Figure reprinted with permission from ACS
Publishers [7])
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 405

Dye + e CB − → reduction products (16.8)

Thermodynamically, water-splitting process (2H2O → 2H2 (g) + O2 (g)) requireds


a Gibbs free energy of 237.18 kJ equivalent to a semiconductor bandgap energy of
1.23  eV (equal to 1.23  V vs normal to hydrogen electrode (NHE)) [7, 10].
Figure 16.2c shows the energetic structure of TiO2 and other semiconductors with
respect to NHE scale. As it can be observed from Fig. 16.2c, bandgap energy of the
semiconductor dictates the water oxidation (2H2O + 4(h+) → O2 + 4H+) and reduc-
tion (2H + 2e− → H2) process [7]. Similarly, to perform the hydroxide (OH•) radical
formation (Eq.  16.2), valence band of the semiconductor must be more positive
potential than O2/H2O. In the case of superoxide (O2−•) radical formation, conduc-
tion band of the semiconductor must be more negative potential with respect to
NHE. The photocatalytic water pollutant oxidation is downhill reaction [11] where
the reactant possesses high energy than the products. Therefore, bandgap energetic
structure of semiconductor plays a crucial role in photocatalytic water pollutant
treatment.

16.3  Analysis of Photocatalytic Performance

In order to ensure the PC reaction in removing pollutants from various water


sources, it is necessary to analyse the source before and after the experiments. The
important parameters which need to be analysed are the amount of pollutants pres-
ent, amount of by-products, chemical oxygen demand (COD) or biological oxygen
demand (BOD).

16.3.1  A
 nalysis of Pollutant Degradation
and By-Product Formation

In general, chromatography, spectrophotometer and chemical analysis techniques


were practised in photocatalytic water treatment analysis. In this session, the prin-
ciples of such analytical tools and how to analyse the resultant data will be discussed.
High-performance liquid chromatography (HPLC)  HPLC is a popular choice of
analytical equipment to identify the complex organic pollutants such as pharmaceu-
ticals, herbicides, surfactants and heavy metals in various water sources. It helps
determine the amount of pollutants in the water and evaluate any intermediate prod-
ucts which can be produced during the photocatalytic degradation process. HPLC
works by injecting the sample into a packed column (either polar or non-polar com-
pounds), while the chosen mobile phase (non-polar or polar based on column) is
passing through the column, which causes separation of the compounds due to the
different interactions between the compounds with the packing material. For the
406 K. R. Davies et al.

analysis of pharmaceuticals and pesticides, the most common type of HPLC used is
reverse phase which involves a non-polar column, while the mobile phase is polar
solvent. The non-polar compounds form attractions with hydrocarbon groups with
the compounds in the sample, thus slowing their travel through the column, while
polar compounds will pass quickly through the column. The detector is usually a
UV light absorption technique as most organic compounds absorb UV light, how-
ever a mass spectrometer can be used too. Prior to HPLC measurements, the known
amount of concentration vs peak area plot has to be derived. An example of HPLC
results and calibration curve is presented in Fig. 16.3. The slope of the linear straight
line as illustrated in Fig. 16.3 will help to quantify the pollutant degradation concen-
tration from HPLC spectra by utlizing the calibration curve:

Concentration = ( A − I ) / m (16.9)

where “A” is the peak area obtained from HPLC, “I” is the y-intercept from the cali-
bration curve and m is the slope of the calibration curve.
The samples were collected periodically at different times during the photoca-
talysis reaction. As explained earlier, the estimated concentration (C) of the pollut-
ant for different times can be obtained using Eq. 16.1. The pollutant degradation
rate can be studied from the following relation:

Fig. 16.3  Typical HPLC results of atazanavir sulphate at 249  nm. Inset: Calibration curve of
(concentration vs peak area) atazanavir sulphate (Figure reused with permission from Elsevier
Publishers [12])
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 407

Percentage of pollutant in sample = ( C × V × D ) / (1000 × W ) (16.10)

where “C” is the concentration (mg/L) of pollutant, “D” is the dilution factor (if
necessary), “V” is the final make-up volume of test solution and “W” is the weight
of sample used for test.
The ratio of pollutant concentration after reaction (C) and before reaction (C0) vs
reaction time plot will explain the pollutant degradation rate. The typical HPLC
results of photocatalytically treated sulfamethazine degradation under solar light is
presented in Fig. 16.4a. From this result, the concentration of pollutant can be esti-
mated from peak area, and the retention time identifies the compound. The plot of
C/Co vs time will provide the order of degradation rate. The typical C/Co plot pre-
sented in Fig. 16.4b is derived from Fig. 16.4a which clearly describes the sulfa-
methazine concentration degradation as time increases.
Example  You started with a solution which contained 150  μg/mL of atazanavir
sulphate. After running the photocatalytic experiment for 2 h, you collected 2 mL of

Fig. 16.4 (a) Sulfamethazine degradation peaks obtained by HPLC mediated by ZnO/CuO)


nanowires under simulated solar light irradiation and (b) effect of light intensity on the sulfametha-
zine degradation kinetics by ZnO/CuO under simulated solar light (Figures reprinted with permis-
sion from MDPI Publishers [13])
408 K. R. Davies et al.

the solution and ran it through the HPLC, and it gave a chromogram which showed
a peak at 8 mins with an area of 800. What is the concentration of the solution and
the percentage degraded?
Using the values obtained from the calibration curve above with the formu-
las, we can:

Concentration = ( 800 − 37.732 ) / ( 23.427 ) = 32.538 µg / mol


Degradation% = ( (150 − 32.538 ) / 150 ) * 100 = 78.31%



The reactor design also influences the efficiency of photocatalysis reaction. It is
important to have a better understanding of the affecting operational factors such as
the initial pollutant concentration and light intensity on the degradation rate. The
first-order kinetic expression has often been used due to its simplicity with good
agreement for a certain initial organic content in photocatalytic processes:

r = ( −dC / dt ) = kapp .C (16.11)

Gas chromatography and mass spectrometry  Gas chromatography coupled with


mass spectrometry is often used for water analysis due to accurate identification of
compounds even at low concentrations in an automatic process. The automation of
the process allows this type of analysis to be fast and easy. The liquid sample is
injected into the inlet port and vaporized to become a gas. The inlet gas helps move
the sample gas through the column where each individual compound reacts with the
column differently resulting in separation of compounds. The individual compounds
leave the column at various times which then get detected by the mass spectrometer
which is used to identify the compounds. The MS utilizes a high beam of electrons
to break the individual compounds into fragments which after acceleration and
deflection hits the detection plate. This calculates the mass to charge ratio (mass
divided by charge) and relative abundance (the amount of fragment present), thus
allowing you to determine the molecule from the MS library. The resulting mass
spectrum is compared with the reference mass spectra by spectrum matching. There
are various factors in which compound identification depends on mass spectrum
library, spectral similarity and weight factors. Retention index from the GC helps
remove potential false-positive identification in the mass spectrum [14].

UV-Vis Spectrophotometer  Some pollutants such as dye and pesticides can be


quantitatively estimated by a spectrophotometer as they can absorb light. The
amount of light absorbed is dependent on the concentration of the sample; this helps
the degradation of pollutants. The UV-visible beam passes through the sample in
cuvette which absorbs the radiation. Utilizing Beer-Lambert Law, we can calculate
the concentration of pollutant from the UV spectra:
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 409

log10 ( I 0 / I ) = ε .l.c (16.12)

which can also be written in accordance to absorbance:

C = ( A ) / ( ε .l ) (16.13)

where, A = absorbance, e = molar absorptivity (found in a table for the chosen wave-
length), l = length of solution that the light passes through (cm) and c = concentra-
tion of solution (mol dm−3).
Example: Backes et al. [15] demonstrated pure TiO2 nanotubes in photocatalytic
degradation of indigo carmine dye. The initial concentration of the dye was 75 μg/L
in water. UV-Vis spectrometer was used to identify the amount of indigo carmine
that was being degraded (please refer to Fig. 16.4).
Utilizing this spectrum, we will be able to find the final concentration of the
indigo carmine pollutant in the solution and the degradation percentage. First step,
we need to identify the absorption quantity at the beginning and end of the experi-
ment. From Fig. 16.5, at the initial concentration, the solution showed an absorption
of 2.18, while at the end had an absorption of 0.18. Using the initial absorption and
concentration with the length of solution in the cell (we assume 1 cm for this experi-
ment), we can calculate the molar absorptivity (which is a constant for that
wavelength).
Convert concentration to molarity:

C (ppm to g/L) = ( 75 * 1000 ) / 1000000 = 0.075 g/L

C (g/L to mol/L) = 0.075 / 466.35 = 0.00016 mol/L

ε = A / l.c = 2.18 / (1 ∗ 0.00016 ) = 13625

Then you can use the same equation with your molar absorptivity value to calcu-
late the final concentration of indigo carmine:

C = A / l.ε = 0.18 / (1 ∗ 13625 ) = 0.000013211mol / L

Finally, we can calculate the degradation percentage as follows:

Degradation percentage = ( 0.00016 − 0.000013211) / ( 0.00016 )  ∗ 100 = 91.74%



Analogy to HPLC, a calibration curve can be established to help determine the
concentration. This avoids the necessity of finding molar absorptivity and dimen-
sions of the cell. From the calibration curve, we are able to find a linear relationship
between the absorbance and concentration, when dilute solutions are used. It is best
to use concentrations which are either side on the concentration you are looking to
find, thus allowing you to find the concentration by looking at the calibration curve.
Chemical oxygen demand (COD)  It is an essential parameter for water analysis
as it measures the oxygen requirement for the oxidization of organic compounds in
water [16]. Thus, it indirectly indicates the amount of organic matter in the contami-
410 K. R. Davies et al.

Fig. 16.5  Photodegradation of an indigo carmine (IC) solution vs time in typical photocatalytic
experiments performed using RF-TiO2 NTs under visible light illumination (cut of filter ≥400 nm).
Note that initial IC concentration was 75 ppm (molar weight 466.35 g/mol). (Figure reprinted with
permission from [15])

nated water. The most common method is known as wet chemistry. It works by
utilizing a chemical oxidant (typically potassium dichromate) to oxidize the
­substance. Therefore, an excess amount of oxidant, set amount of sulfuric acid and
heat (usually 150), is added to sample for digestion to take place. While metal slats
can be utilized to catalyse the digestion and suppress any interferences, the titrimet-
ric or colorimeter is utilized to measure the excessive oxidant. The titrimetric
method utilizes ferrous ammonium sulphate as a reducing agent which converts the
excess dichromate into its trivalent form. The point in which all the excess dichro-
mate has converted allows calculating the amount of dichromate in the sample as
it’s equal to the amount of ferrous ammonium. By subtracting this amount from the
amount originally added to the sample results the amount of dichromate that went
toward oxidizing the contaminants, while colorimetric method utilizes the absor-
bance capabilities of the sample. The amount of trivalent chromium present in the
sample after digestion can be determined by measuring the absorbance of the sam-
ple at 600 nm in a photometer or spectrometer. Both methods have their advantages
and limitations, for example, titration requires less equipment but requires more
time and work, while the colorimetric is more expensive for instruments; however
less work is required as you can leave the samples in the instruments to digest and
do the work.

Total organic carbon (TOC)  This measurement is a parameter which is widely


found in wastewater treatment to determine the amount of carbon in organic com-
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 411

pounds in the water [17, 18]. TOC varies in each water source due to the presence
of dissolved organic carbon, non-dissolved organic carbon, non-organic carbon,
total carbon, volatile organic carbon and non-purgeable organic carbon. To deter-
mine the TOC of the water source, then the quantification of the products is formed
by the oxidization of the organic compounds. The oxidation of these organic com-
pounds can be done by several techniques depending on the users [19, 20].

16.4  Pollutants

There are many pollutants currently affecting freshwater sources and thus impacting
on our drinking water. This is due to existing water technologies being inadequate
(reverse osmosis, nanofilteration, coagulation and sedimentation) and unable to
remove 100% of the pollutants. In addition, these water treatment technologies
either require high energy demand or high costs or waste production. To improve
the water treatment technologies, it’s necessary to understand the pollutant’s behav-
iour in the water.

16.4.1  Organic Dye Pollutant

Synthetic dyes utilized in the textile industry contain extremely toxic chemical
properties that affect water quality while it discharges into waterbodies. Dyeing is a
big envinromental issue as it supplies 15–20% of the total wastewater flow [21]. For
instance, the textile industry produces over 700,000 tons of synthetic dyes while
utilizing 40,000 different dyes and pigments [22, 23]. This has resulted in 20,000
tons of dye going into the wastewater every year due to ineffective cleaning tech-
nologies [23] as a result from the complex aromatic structure. This means that the
toxic pollutants stay in the environment for a long period of time as Hoe et al. dem-
onstrated with Reactive Blue 19 which has a half-life of 46 years [24]. Since 1997
the environmental policy in Britain has made it illegal for industries to release syn-
thetic chemicals into any water sources. This is a massive challenge for the textile
industry as many conventional treatment systems cannot fully remove these dyes
due to their chemical properties. Due to this environmental policy, many textile
industries have relocated to developing countries that have less stringent policies.
This is not a solution as these dyes are now affecting many people in the developing
countries. The untreated or partially treated dye pollutants have numerous side
effects which can adversely affects the environment and public health due to their
chemical properties [25]. One example is azo dyes (a molecule that has two adjacent
nitrogen atoms between carbon atoms) which have carcinogenic properties that
cause organisms (humans included) to develop cancer when consumed or exposed
to the dye depending on the period of time and concentration of the dye [26]. Even
at very low concentrations (<1 ppm), azo dyes can cause issues with marine life as
412 K. R. Davies et al.

their presence in water increases the turbidity which prevents light penetration, thus
reducing photosynthesis activity and creating an oxygen deficit [27]. Dye-
contaminated water can also cause issues in farming as this type of pollutant blocks
the pores in the soil, thus resulting in the soil becoming hard and the crops’ roots not
being able to attach to the soil, causing loss in production [28].

Advances in Photocatalytic Dye Degradation

Due to its high photocatalytic activity and excellent biocompatibility property, the
most popular semiconductor photocatalyst in dye degradation is TiO2 [29]. TiO2 has
shown excellent photocatalytic degradation performance in organic pollutants relat-
ing to nontoxic substances such as CO2 and HCl [30]. However, the efficiency of
this photocatalyst is based on its operating conditions such as pH, type of dye, pho-
tocatalyst concentration and oxidizing agent. Khataee et al. reviewed that dyes with
an anthraquinone structure had a lower degradation rate than monoazo dyes [31].
The molecule groups present in the dye is a crucial factor, for example, methyl and
chloro groups decreased the removal efficiency, while nitrite increases it [31]. The
hydroxyl groups favour degradation due to the heightened electron resonance, while
having more sulfonic substituents make the dye less reactive in photocatalysis [31].
The solubility of the dye can also decrease due to alkyl side chains [31]. Therefore
it is essential to understand the dye molecule structure when adopting photocataly-
sis technique [31].
Natural sunlight is green, free and abundant. Therefore, sunlight-driven photoca-
talysis will be a sustainable route compared to conventional water treatment tech-
niques. S. Sakthivel et al. [32] explored TiO2 supported on alumina and glass beads
which completely mineralized acid brown 14 after 420  min in natural sunlight.
However, due to the varying structures of dyes, the same photocatalyst system does
not work as effectively for all dyes; hence modifying the semiconductor structure
helps improve the photocatalysis rate.
The major challenge of utilizing TiO2 is its wide bandgap energy, thus optical
absorption limited to the UV light [33]. Therefore, there has been research pro-
gressed on alternative TiO2 or modified TiO2 photocatalysts. In 2019, Na Quin et al.
found that utilizing CuS (3D hierarchical network structure) helped remove 100%
of methylene blue (MB) and rhodamine B (RhB) (individually) when irradiated by
25 min [34]. It was also showed to be effective when both dyes were present in the
solution as it removed 99% of MB and RhB when utilizing 0.5 g of CuS, 25 min of
solar light irradiation, 40 mL (10 mg/L) of MB/RhB solution, 40 μL of H2O2, pH of
4 and 30 °C [34]. This shows that CuS is an effective photocatalyst and stable pho-
tocatalyst (no drop in degradation when the experiment was repeated four times)
which can be utilized for dye wastewater which contains multiple dyes [34]. Chen
et al. utilized ZnO as a photocatalyst to improve dye degradation [35]. They devel-
oped a ZnO photocatalyst with a composite ratio of 4:1 and calcination temp of 400;
this removed 99.70, 97.53 and 89.59% of MO, CR and DB38, respectively, from the
water solution. Even after four cycles, it showed only a minimal decrease in
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 413

degradation due to loss of the photocatalyst. To improve the degradation, then


increasing the dosage of catalyst and decreasing the initial concentration of azo dye
are recommended. The reaction was favoured in acidic conditions due to the type of
dye [35].
Many researchers have demonstrated doped TiO2 with a metal and non-metal
species (second-generation photocatalysts) in order to narrow the bandgap while
also forming a heterojunction. This has been proven to show superior dye degrada-
tion when compared to first-generation photocatalysts (TiO2, ZnO, WO3), as the
heterojunction helps suppress the recombination of photogeneration charge carriers.
Ce-doped ZnO (supported on chitosan chains) has been demonstrated by Saad et al.
[36]. The electrons from the CB of the ZnO move into the CB of the Ce due to the
CB of Ce being a lower energy than the CB of ZnO, while the holes remain in the
valance band (VB) of the ZnO, thus preventing any recombination of chargers and
allowing higher degradation. It showed that Ce-ZnO degraded the malachite green
dye (MG) 33% more than pure ZnO (supported on chitosan chains), due to this
prevention of charge carriers while also lowering the bandgap to 2.5 V which allows
more wavelengths to be adsorbed [36].
The doping effect widens the absorption range from UV to visible wavelength,
increasing the carrier charge separation, thus improving the removal rates. Chi Hieu
et al. [37] reported palladium (Pd)-doped TiO2 more effectively degrade MB and
MO dye under UV irradiation than undoped TiO2. They found that in neutral pH a
photocatalyst with 0.5  wt% of Pd-doping, the MB solution was decolourized by
96.9–99.4%, while the MO solution was decolourized by 82.5–92.6% [37]. The
mineralization however was lower, 85.9% and 77.1% for MB and MO, respectively;
this was due to the by-products produced causing competition for absorption [37].
However when the two dyes were mixed in the same solution, the optimum doping
increased to 0.75% as the presence of another dye caused increased competition for
active sites, thus decreasing the removal rate [37]. Abdelaal et al. [38] utilized an
Pd/TiO2-CS photocatalyst to remove methylene blue from water under visible light.
They found that the optimized conditions were 0.75 g of the catalyst in 500 mL of
methylene blue solution (concentration 75 mg/L) which degraded 99.5% of methy-
lene blue within 30 min (Fig. 16.6a) of visible light irradiation, repeating the experi-
ment with the same catalyst found effective after six cycles (Fig. 16.6b) [38]. This
clearly shows that the doping technique has many advantages such as lowering the
recombination rate, and the addition of CS helped prevent agglomeration, thus
improving the degradation of the organic dyes [38].
Recovery of powder-type photocatalyst after the dye degradation reaction is a
challenging task. From an industry point of view, it is necessary to be able to reduce
costs, such as removing a post treatment to eliminate nanosized photocatalysts in
suspensions. One option is immobilizing the photocatalyst onto a support. For
instance, Kuo et al. immobilized TiO2 onto the inner surface of the reaction vessel
which achieved 100% colour removal for methylene blue (5 μM) after 6 h of solar
light [39]. For UV light irradiation, the colour removal was lower; this suggested
that the increased temperature from the solar light helped promote decolouring. It
414 K. R. Davies et al.

Fig. 16.6 (a) The degradation of methylene blue by utilizing different concentrations of Pd/
TiO2-CS photocatalyst. (b) The photocatalytic degradation percentage for every recycle of the
photocatalyst. (Figure is reprinted with permission from Elsevier Publishers [38])

was identified that to increase the colour removal, then an acidic pH and lower ini-
tial concentration should be utilized [39].
The pH of the solution medium modifies the electrical double layer of the solid
electrolyte interface and consequently affects the sorption-desorption processes
[40]. This indirectly influences dye degradation rate through affecting charge sepa-
ration of the photogenerated electron-hole pairs in the surface of the semiconductor
particles [40]. Interestingly, TiO2 exhibits an amphoteric character which can
develop either a positive or a negative charge on its surface [41, 42]. Ling et al. [43]
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 415

identified that under basic pH condition the electrostatic interactions caused strong
absorption between the negative TiO2 and methylene blue cations. While Baran
et  al. [44] found that acidic conditions increased the degradation of bromocresol
purple due to positive charges of the dye. Thus proving the effect of the pH is depen-
dent on the target dye charges, with cationic dye degradation being favoured in the
alkaline conditions, and anionic dye degradation is favoured in acidic conditions.
The ratio of pollutant concentration and photocatalyst solid quantity is a key
processing parameter to determine the pollution degradation efficiency. For instance,
increasing the initial concentration of the water pollutants causes the pollutant deg-
radation to lower due to two reasons. One being that increasing the initial concentra-
tion means the ratio between pollutants to activate sites lower thus meaning there is
less readily available sites compared to the amount of pollutants present. This was
demonstrated by Saad et al. [36] who found increasing the initial concentration of
MG from 5 mg/L to 15 mg/L lowered the degradation rate as it took 90 min for
complete removal at 5 mg/L, while for 15 mg/L complete degradation was achieved
in 210 min. Also they demonstrate that irritation time is essential in achieving good
removal as increasing the time increases the degradation as more photons are
adsorbed; thus more electron-hole pairs are produced [36]. In view of optical path
length at semiconductor photocatalyst increasing the pollutants, concentration dye
can block light from penetrating the photocatalyst particles, thus lowering the num-
ber of electrons being excited and reducing the dye degradation efficiency. This
indicates that photocatalysis could be a final polishing step in order to remove dyes
from water which contain low concentration of dyes. However, Waghmode et al.
[45] found that utilizing photocatalysis as a pre-step process before a biological
treatment was able to achieve 100% decolourization and 80% COD removal which
is higher than when only biological treatment was utilized while also requiring a
shirt process time. This could be an option for industries to use in order to achieve
better dye degradation while reducing their treatment time (thus lower cost).
Therefore, to improve dye degradation then increasing the catalyst dosage in the
solution is an option so that the number of active sites and the adsorption capacities
of the catalyst increases. This means less irradiation time is required as treatment is
quicker or higher concentration of dye could be treated.

16.4.2  Pesticides

Pesticides are any chemical compound that can kill or control pests. According to a
US Environmental Protection Agency (EPA) report in 2015, approximately 5.2 bil-
lion pounds worth of pesticides were being used [46]. The most popular pesticides
used are insecticides (approximately 44% of all pesticides) which kill both plants
(such as weeds) and insects (such as slugs). Pesticides are known to be toxic after
one ingestion, inhalation or skin contact. This can cause respiratory tract, eye and
skin irritation, allergic reaction, nausea, vomiting and diarrhoea, headaches and diz-
ziness, extreme weakness and seizures [47]. When continuously exposed to these
416 K. R. Davies et al.

chemicals, the effects become more harmful even at low concentrations [47]. These
effects include Parkinson’s disease, asthma, depression, infertility, cancer, attention
deficit and hyperactivity disorder and death [48]. In addition these pesticides com-
bine with other chemicals in the water which can increase these effects and cause
even more issues [49].
The TiO2 photocatalyst has been tested in several pesticides including isopro-
turon, alachlor, diuron, atrazine in ultrapure and natural water [50]. Marta Cruz
et al. [51] used 500 mg L−1 of TiO2 to achieve nearly 100% of pesticide removal in
ultrapure water after 300 min of UV-A irradiation. However for natural water, the
removal of pesticides was significantly lower [51]. This is due to natural water con-
taining dissolved organic and inorganic contaminants which increase competition
with hydroxyl radicals, resulting in lower degradation of pollutants [51].
In order to improve the pesticide degradation using a TiO2 photocatalyst, some
researchers have adopted a doping strategy as explained earlier. Marta Cruz et al.
proved that graphene oxide (GO) doping with TiO2 improved photocatalysis perfor-
mance at visible light due to lower bandgap allowing a greater light absorption [51].
Here, GO helps facilitate charge separation as the electrons move from the TiO2 into
the GO, thus improving the degradation [51]. Another doping strategy was using
rare earth elements to help increase the wavelength adsorption range into the visible
light and enhance the photocatalytic activity. Doping with rare elements help
improve the textural, structural, electronic and photocatalytic properties of TiO2.
Recent testing of degradation by photocatalyst on diuron [52, 53] and methyl para-
thion pesticide [54, 55] degradation by photocatalysis is noteworthy to consider as
proof-of-concept type of demonstration. The rare earth element-doped TiO2 exhibits
improved photocatalytic pesticide degradation under optimum condition [56]. Pérez
et  al. demonstrated the La-, Ce- and Sm-doped TiO2  in photocatalytic pesticide
degradation. The experiment contained 0.5 g/L of catalyst in 300 mL of solution and
a flow of 60  mL/seg atmospheric O2 which was carried out under sunlight for
300  min [57]. The TiO2 doping with La and Ce (0.1% and calcination of 500C)
showed nearly 100% degradation of diuron, due to the increasing volume of pores
in TiO2 surface achieving greater contact between solution and photocatalyst.
However, some rare earth elements, for example, Eu, have the opposite effect as
they block the pores. Increasing the dopant to 0.3% only increases the photoactivity
performance of La, Ce and Sm; however increasing to 0.5% causes a decrease in
photocatalytic efficiency as the pore diameter and crystal size decreases and allows
recombination to occur. While for methyl parathion EU 0.3% doped TiO2 demon-
strated the best degradation of above 95%, with the other rare elements only achiev-
ing above 70% for degradation of methyl parathion [57]. This shows that degradation
of the pesticide is dependent on the dopant and the calcination temperature as it
varies the morphology of the catalyst which can either increase the contact or
decrease the contact with solution as well as suppressing the e−/h+ recombina-
tion [57].
It is well recognized that immobilization helps photocatalyst to be recovered
after reaction and is appropriate for industrial application. However, it needs to be
optimized to ensure there is enough contact with the catalyst and pollutants. Olga
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 417

Sacco et al. [58] investigated the photocatalytic removal of atrazine using N-doped
TiO2 supported on ZnS-based phosphors microparticles, thus creating a core-shell
catalyst. They found 94% of atrazine was removed after 90 min of UVA irradiation
using N-TiO2/ZPS (30 wt% loading of N-TiO2) when using 0.5 g L−1 photocatalyst
at pH of 5.8 [58]. The presence of ZPS was important and increased the degradation
when compared to pristine N-doped TiO2 due to its intrinsic photoactivity [58].
Thus it is important to select a good support to help increase degradation. Another
immobilization technique was investigated by C. Belver [59] who utilized W-doped
TiO2 anchored on clay for the removal of atrazine, with the best photoactivity being
exhibited by W-TiO2/clay catalyst with 5 wt% of W in solar light irradiation. The
doping of W narrows the bandgap of TiO2, thus allowing the absorption of visible
light and improving the degradation of atrazine in solar light while having a better
charge separation of e−/h+ [59].
There have been some recent advances in the use of photocatalysis using real-life
wastewater containing pesticides. A real-life sample was investigated by Miguel
et al. [60] to determine if photocatalysis could be used for Ebro River Basin (Spain)
which contained 44 organic pesticides. Just after 30 min of light irradiation resulted
in the average pesticide having a 48% degradation when utilizing 1 g L−1 of TiO2
while increasing to 57% when (10 mM) hydrogen peroxide was added [60]. The
highest degradation was 80% which is achieved by parathion methyl, α-endosulfan,
chlorpyrifos, 3,4-dichloroaniline, 4-isopropylaniline and dicofol, while some pesti-
cides achieve less than 30% (HCHs, endosulfan-sulphate, heptachlors epoxides and
4,4′-dichlorobenzophenone) [60]. Sewage wastewater effluent (100 L) containing
six pesticides with endocrine disrupting activity was cleaned at pilot plant scale
under natural sunlight in Murcia, SE, Spain [61]. The reactor contained five boro-
silicate tubes mounted on curved polished aluminium reflectors with water flow
provided by centrifugal pump, with the system being continuously stirred [61]. The
product water was filtered through an ultrafiltration membrane to recover the TiO2
photocatalyst. Initially 0.30 mg L−1 of each pesticide in 100 L of wastewater were
tested with two commercial TiO2 photocatalyst powders (P25 and Kronos vlp 7000)
[61]. Later, this experiment was upscaled using 200  mg  L−1 with the addition of
250 mg L−1 of Na2S2O8 [61]. It resulted in 75% (±25% depending on the pesticide)
of pesticide to be degraded. When utilizing a P25 TiO2 photocatalyst, malathion was
totally degraded, while fenarimol was more resistant (55% remaining). This was an
increase compared to utilizing a TiO2 vlp 7000 photocatalyst due to the structure of
the TiO2 (Fig. 16.7). The P25 TiO2 has a structure of 70/30 of anatase/rutile, respec-
tively, while TiO2 vlp 7000 has a structure of entirely anatase [61]. Therefore, due
to the rutile and anatase heterostructure in the P25 photocatalyst there’s an improved
charge separation through effective electrons transferring from rutile to anatase,
thus suppressing the e/h recombination and improving photodegradation perfor-
mance [61].
418 K. R. Davies et al.

16.4.3  Pharmaceutical Pollutants

Many studies have found that the concentration of pharmaceutical pollutants are
low in sewage effluents and surface water; however these studies do not include
drug manufacturing sites [62]. When Larsson et al. [62] conducted studies in the
effluent discharge from Patancheru Enviro Tech Ltd, India (wastewater treatment
plant which receives effluent from 90 drug manufacturers), they found high levels
of several drugs. The issue with pharmaceutical drugs entering the water system is
that their parent compounds cannot be eradicated by current water treatment

Fig. 16.7  Change of pesticide concentrations as a function of natural sunlight irradiation time
using the tandem TiO2 (200 mg L−1)/Na2S2O8 (250 mg L−1) at pilot plant scale. (Figure reprinted
with permission from Elsevier Publishers [61])
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 419

methods [62]. The low concentration (within the allowed limit) of the individual
drug would not pose a threat to human health; however the combined concentration
of all the pharmaceutical drugs released would affect organisms especially over an
extended period of time [62]. A few more studies found that in the USA, Germany
and the Netherlands, the final drinking detected trace levels of low concentrations of
pharmaceutical drugs [63]. There are many ways in which pharmaceutical com-
pounds reach the water supply: incorrect disposal of pharmaceutical waste, excess
medication consumption, and people or animals excreting medication through uri-
nation or vomiting [62, 63]. Human and animal excretion of pharmaceuticals is a
major issue as it releases pharmaceutical metabolite which can be as harmful as
their parent compound (refer to Service, B. I. Study on the environmental risks of
medicinal products; Final Report prepared for Executive Agency for Health and
Consumers, 2013).
A recent study by Simsek et al. [64] identified that boron-doped TiO2 (B-TiO2)
utilizing the solvothermal method has improved the degradation of pharmaceuticals
(ibuprofen and flurbiprofen) in visible light irradiation. The B-TiO2 showed an
improved degradation of the pharmaceutical compounds 81–85% after 5  h com-
pared to undoped TiO2 (40%) due to several reasons such as (a) the decrease in
bandgap energy and (b) boron acting as an electron trap which decreases the recom-
bination rate, thus allowing more holes for reduction of pharmaceuticals [64]. This
experiment was conducted with 100  mL of solution with the pollutants being
20 mg L−1, while the catalyst was 1 g/L. Mohamed Gar Alalm et al. [65] demon-
strated the photocatalytic degradation of amoxicillin, ampicillin, diclofenac and
paracetamol by immobilizing TiO2 onto activated carbon (AC) from water under
solar irradiation (Fig. 16.8). In this study, TiO2/AC was able to completely degrade
amoxicillin and ampicillin, while bare TiO2 only degraded by 89% and 83%, respec-
tively [65]. In the case of diclofenac and paracetamol, the TiO2/AC photocatalyst
degraded around 85% and 70%, respectively, which was more effective than pure
TiO2. The optimum conditions were found to be 1.2 g/L of TiO2/AC, while 0.6 g/L
for TiO2. It is noted that beyond this quantity too much turbidity was created due to
the fine particles [65]. Interestingly, as estimated total costs for usage of immobi-
lized as well as pure TiO2, the TiO2/AC is to be cheaper even at the optimum condi-
tions of (400  mg/L of TiO2 and 800  AC dose), thus proving immobilization can
improve the photocatalytic activity and cost-effective industrial applications [65].
Yujie He [66] found that immobilizing TiO2 on sand (200–500 micron) resulted in
high removal of pharmaceutically active compounds (PhACs) in wastewater efflu-
ent. The PhACs were removed propranolol (100%), diclofenac (100%) and carbam-
azepine (76%) after 96 h of irradiation [66]. This showed that post-treatment process
could be effective prior to discharging.
Mohamed Gar Alalm et al. [65] also found that the pH of the solution influences
the photocatalytic degradation rate due to modifying the photocatalyst and pollutant
surface charge; thus the optimized pH condition requires to be found through exper-
iments. From Fig. 16.8, amoxicillin and ampicillin were degraded 100% at pH 10,
but for diclofenac a pH of 7 achieved high degradation [65]. They found that the
420 K. R. Davies et al.

100
(a) pH 3
80
pH 5
60 pH 7
C/C0 %

pH 10
40

20

100
(b)
80

60
C/C0 %

40

20

100
(c)
80

60
C/C0 %

40

20

100

(d) 80

60
C/C0 %

40

20
Adsorption Photocatalytic
degradation
0
–30 0 30 60 90 120 150 180

t30,w (min)

Fig. 16.8  Effect of initial pH on degradation of pharmaceuticals by photocatalysis. Note that ini-
tial concentration of all pharmaceuticals = 50 mg/L, TiO2/AC dosage = 1.2 g/L. (a) Amoxicillin.
(b) Ampicillin. (c) Diclofenac. (d) Paracetamol. (Figure reprinted with permission from Elsevier
Publishers [65])
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 421

removal efficiency for amoxicillin and ampicillin was lower at pH values ranging
from 3 to 5 [65].
To explain the effect of pH on PC, then ionization states of TiO2 and dopant need
to be considered. The zero point of charge (pHPZC) and pH of solution affect the
surface charge of the photocatalyst [65]. TiO2 has been documented to have a pHPZC
of 6.4 which changes from a positive surface charge (pKa1  =  2.6) to negative
(pKa2 = 9) when the pH increases [67]. For amoxicillin the charge is positive in
acidic pH, while negative in alkaline pH.  Thus in acidic conditions, amoxicillin
repels from the photocatalyst due to both being positively charged, which is a simi-
lar case in alkaline solution as they are both negatively charged. However, alkaline
solutions help increase degradation of pollutants (please refer to Fig. 16.8) due to
the higher production of hydroxyl radicals from increased availability of hydroxyl
ions present on the TiO2 surface [67]. It implies that the solution should be kept in
neutral to alkaline conditions as acidic conditions cause repulsion and do not have
the hydroxyl ions available for hydroxyl radical production (Fig. 16.9).
Recently Jallouli et  al. [68] demonstrated TiO2/UV-LED system in degrading
ibuprofen (IBU) from pharmaceutical wastewater. It was found that IBU was
reduced to below the detection limit when utilizing TiO2 heterogeneous photocata-
lyst with UV-LED light source. While in secondary treated effluent (collected in
wastewater treatment plant), the IBU degradation was lower due to the competition
with other species dissolved in the water. The optimum pH for the treatment of pro-
cess industry wastewater (PIWW) was found to be 5–5.3 as it allowed ibuprofen
(point of zero charge  =  4.9) to be negatively charge, while TiO2 (point of zero
charge  =  6.4) would be positively charged which improves the electrostatic

Fig. 16.9  The degradation of ibuprofen when utilizing different numbers of LEDs for TiO2 photo-
catalysis. (Figure reprinted with permission from Elsevier Publishers [68])
422 K. R. Davies et al.

attractions, thus increasing the degradation of IBU [68]. They found that utilizing
four LEDs resulted in 100% degradation of IBU from PIWW after 30 min of irra-
diation when using 1 g L−1 of catalyst in pH 5.3 which is higher than two LEDs [68].
This is because the high light intensity allows more photons to be adsorbed, thus
producing more hydroxyl radicals from more electrons transferring. The effect of
catalyst loading showed that 2.5 g L−1 of TiO2 achieved the highest degradation and
mineralization (57%) [68]. This evidences the potential for photocatalysis to be
used in the industry for parametrical removal.
It is important to not only analyse the amount of pharmaceutical degradation but
also determine the by- and intermediate products. This means a mass spectrometer
needs to be utilized in order to determine what intermediates are being produced
during the process, while COD and TOC help determine if these products are being
fully degraded. While this works for known pollutants in pure water, a toxicity
assessment is required for real wastewater treated by photocatalysis.
When treating real-life wastewater samples, it is essential to determine the toxic-
ity of the treated water to ensure that water returning to the environment is not toxic
to the aquatic species or cause any risk to animals in the higher levels of the food
chain. To determine toxicity of the water, then either bacteria assessments such as
Aliivibrio fischeri and E. coli (the Kirby-Bauer method) or aquatic invertebrates
such as Crustaceans and Daphnia magna are widely utilized [69, 70]. Even though
bacteria are less effected by toxicity of nanoparticles, they have a short lifespan and
easy to cultivate, thus making them an easy option for toxicity assessment [70].
While non-complete mineralization of pharmaceutical pollutants means there is still
toxic compounds in the wastewater, the utilization of photocatalyst nanoparticles in
powder form can cause toxicity [69]. This is due to the leakage of nanoparticles into
treated water due to filtration not completely removing all the nanoparticles from
the treated water. There are reviews which highlights work which has shown the
toxicity effects of nanoparticles and why. One study conducted by Blaise et al. [71]
showed the presence of 1 mg/L TiO2 nanoparticles caused oxidative stress in the
gills of rainbow trout (Oncorhynchus mykiss) after 14 days of exposure, thus show-
ing even a small amount of TiO2 can cause bad effects on the aquatic life. This type
of nanoparticles greatly affects the toxicity such as wastewater treated by TiO2
nanoparticles and has shown to lower the toxicity when compared to ZnO due to the
higher effectiveness of TiO2 and lower dissolution occurring, and silver nano-plates
demonstrated higher toxicity to zebrafish embryos than wires and spheres [69, 70].
However, this is not the only concern, but also the type of pollutant absorbed onto
the nanoparticle can greatly affect the toxicity of the treated water such as TiO2 used
for olive oil mill wastewater that caused toxic effects for Chironomus riparius (tox-
icity assessment), while TiO2 used for mine draining did not have any negative
effects on Chironomus riparius [69]. This shows that for industrial use it’s key to
avoid using nanoparticles as powder form but instead to immobilize them onto a
substrate in order to minimize the chances of leakage into the treated water while
avoiding the need to have a post-filtration step [69]. However, there is currently a
limited amount of studies utilizing immobilised photocatalysts in a photocatalysis
process due to its low effectiveness and higher production cost. Therefore, it might
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 423

be recommended that photocatalysis is utilized as a pretreatment step in order to


convert the pharmaceutical compounds into more biodegradable products so they
can be fully removed by a biological treatment process and that immobilized photo-
catalysts are used in order to remove any issues with nanoparticle leakage. However,
it’s highlighted in the critical review on application of photocatalysis for toxicity
reduction of real wastewaters that there are very limited comparisons of the effects
of different nanoparticles in wastewater [69].

16.4.4  Heavy Metal Pollution

Photocatalysis has seen some success in isolating pollutants such as soluble heavy
metals (chromium, copper, lead, zinc, arsenic, cadmium, mercury, etc.) in industrial
wastewater [72–74]. These heavy metals are introduced by all manners of industry
such as mining, smelting or electroplating. The removal of unwanted elements from
wastewater is strongly desirable for both economic and domestic reasons, for reuse
in industry or for safe consumption by human or domestic animal populations.
Traditional wastewater treatment techniques are effective at removing soluble pol-
lutants but often result in hazardous by-products such as highly concentrated slud-
ges. The issue here is that, while we can already remove pollutants from wastewater,
the currently utilized methods generate environmentally dangerous waste which is
difficult and expensive to dispose of, not to mention that any potentially reusable
material is suspended in the waste sludge. The process of photocatalysis is often
paired with adsorption techniques to effectively capture and filter pollutants follow-
ing oxidation; some recent techniques are discussed later.
Mostly metal oxidation or reduction is occurred by successive one-electron path-
ways from unstable intermediate to the most stable species formation. In photocata-
lytic metal reduction process, the photoelectron from conduction band of the
semiconductor drives the reaction either directly or indirectly. The electron donors
facilitate the indirect metal reduction process. The photohole or hydroxyl radicals
generated from VB of the semiconductor led to the oxidative removal of the metals
[75]. Figure 16.10a shows the bandgap energetic structure of TiO2 and position of
the reduction potentials of various metallic couples related to the energy levels [75].
This explains the influence of bandgap energy on photocatalytic metal pollution
reduction.
The adsorption material-integrated photocatalyst is showing higher performance
in metal removal than photocatalyst alone [77]. Li et  al. [76] demonstrated the
method of removing Cr (VI) from wastewater using a TiO2-graphene hydrogel,
dehydrated and suspended in wastewater prior to UV activation (Fig. 16.10b).
The benefits of this method include the stability of TiO2 and the 3D network
structure provided by graphene, leading to excellent photocatalysis and adsorption
of Cr(VI). The proposed mechanism of photocatalysis and adsorption follows;
Cr(VI) is adsorbed by graphene hydrogel through the 3D structure of TiO2-graphene
hydrogel. Next the Cr(VI) was photocatalytically reduced by the TiO2, facilitated by
424 K. R. Davies et al.

Fig. 16.10 (a) Bandgap energetic structure of TiO2-P25 compared with reduction potentials of
various metallic couples (Figure reprinted with permission from Elsevier Publishers [75]). (b)
Schematic diagram of TiO2-rGH-mediated removal of Cr(VI) (Figure reprinted with permission
from Elsevier Publishers [76])

close contact to the photocatalyst and the UV e−/h+ activation. The reaction can be
summarized as:

TiO2 + hv → h + VB + e − CB

Cr2 O7 2 − + H 2 + + e − → Cr 3+ + H 2 O

H 2 O + h + → O2 + H +

The Cr(VI) was removed 100% from a solution of 5 mg/L Cr(VI) concentration.
This is quite an accomplishment; however it must be acknowledged that this method
requires UV irradiation and post treatment of the nanomaterial for continued effec-
tiveness. These drawbacks are inherent to TiO2 nanomaterials and should be viewed
accordingly. Heterostructure formation through assembling narrow bandgap mate-
rial BiWO4 onto TiO2 is a recent good example for enhancing the light-driven metal
reduction process [78]. An alternative TiO2 visible light-driven material is also a
promising photocatalyst to demonstrate the heavy metal reduction process.

16.5  Future Outlook

A major limitation to TiO2-based photocatalysis is the low photoconversion effi-


ciency due to each stage of the process having a loss of efficiency [79]. Research has
demonstrated that the quantum efficiency varies depending on the photocatalyst and
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 425

experimental conditions. Photocatalysis has only achieved a small percentage of


quantum efficiency (<10%) due to fast recombination of the photogenerated e−/h+
pairs [1]. Therefore, appropriate doping can help increase quantum yield by reduc-
ing recombination rate. All AOPs have a limitation in which only a small amount of
the generated ROS is involved in the degradation of pollutants (for real-life sam-
ples). This is due to the background compounds such as natural organic matter
being ROS scavengers or absorbing light resulting in a lower efficiency. The TiO2
photocatalysts have an additional constraint, namely, limited light absorption as
well as high recombination rate of photogenerated carriers [5]. Many researchers
have reduced the recombination rate through doping or co-doping which has shown
to increase photodegradation of pollutants. However high amount of co-catalyst has
led to a fragile photocatalyst, with limited surface area and complicated production
of photocatalyst [79].
The design of the photocatalytic reactor changes the efficiency of the photoca-
talysis for the upscaling of the process to the industrial scale. This is especially true
in the case of electrode-type photocatalyst. Powder photocatalysts increases the
photoactivity due to the increase surface area and contact time; however this does
require a post-filtration step which can cause shear stress on the catalyst particles
and increased cost [6]. Another issue associated with powder photocatalysts is the
potential of leakage which can affect the environment [80]. However, there is an
option to immobilize powder photocatalyst onto a substrate, thus avoiding the post-­
filtration step; this does limit their efficiency due to their complicated surface area
which is reduced for each volume unit of water [6].
Another limitation includes the fact that there is no one method for evaluating the
performance of photocatalyst or reactor design due to the high number of various
parameters which affect the performance [79]. This reduces effective comparison of
various photocatalysts/reactor designs to find the most efficient system. There have
been some options such as finding the quantum efficiency through the destruction of
target pollutants or generation of oxidizing species such as OH radicals [79]. As we
have seen in Sect. 16.4.2, the research recently has included real-life samples which
inform the decision on whether this technology can be used by the industry.
However, while it has been tested, it requires more work to help bring this sustain-
able alternative into the industry. Future work should be focusing on trying to find a
photocatalyst which will effectively remove all target pollutants while utilizing an
all solar spectrum. This could be done through doping, dye sensitizing or even find-
ing an alternative photocatalyst. It would be beneficial to the researching society if
there was a database for all current photocatalysts and the results they achieved for
each pollutant tested. This would help reduce time wasted in repeating tests and
could help optimize systems. It is also essential that more research is carried out to
ensure the process is a continuous flow system which achieves low operating and
maintenance costs. It would also be beneficial to establish how long a photocatalyst
can last with constant solar irradiation, for example, if it can work for a month or
longer without changing.
The use of hybrid photocatalysts and adsorbents to treat heavy metal pollution in
wastewater has progressed significantly during the last decade. Foundations for
426 K. R. Davies et al.

further study have been laid and the literature reflects this. The gradual awakening
of governments and general populations to the concerns of heavy metal water pol-
lution have quickened since the turn of the twenty-first century. The immediate
effect of this shifts in research goals to counter large-scale pollution, with funding
to complement. New nanomaterials and synthesis methods are being developed rap-
idly, while well-known ones such as TiO2 undergo hybridization and innovation. In
short, photocatalysis is now understood to be a viable solution to heavy metal waste-
water pollution. As such the research community has delivered an energized
response.

16.6  Conclusion

Photocatalysis has been able to eliminate a large range of pollutants while utilizing
solar energy, thus making it a renewable energy. It has many benefits which make it
a better solution for water technologies. For example, it can totally mineralize pol-
lutants, using a low-cost material, while being nontoxic and producing hydrogen (a
fuel alternative). Though there is a considerable research relating to photocatalysis,
it is still not being fully utilized in the industry due to the issues with upscaling and
most research being conducted in the lab-scale environment. Therefore, it is neces-
sary to carry out research on a larger scale to establish if the process can be achieved
and optimized on an industrial scale with industrial waste. In addition, with a sig-
nificant amount of researchers conducting photocatalytic experiments everyday it
would be advantageous to develop a database which would include the results from
each photocatalyst tested. This would help prevent any duplication of tests and
improve the optimization of the material and conditions of the experiment.

Acknowledgements  S.P. thanks to Sêr Cymru II—Rising Star Fellowship program for support-
ing this work through the Welsh government and European Regional Development Fund (80761-­
SU-­102 (West)). C.T. and S.P. thank Royal Society, UK (IEC\R3\170085-International Exchanges
2017 Cost Share) for partially supporting this work.

References

1. Friedmann, D., Mendive, C., & Bahnemann, D. (2010). TiO2 for water treatment: Parameters
affecting the kinetics and mechanisms of photocatalysis. Applied Catalysis B: Environmental,
99, 398–406. https://doi.org/10.1016/j.apcatb.2010.05.014.
2. Fujishima, A., & Honda, K. (1972). Electrochemical photolysis of water at a semiconductor
electrode. Nature, 238, 37–38. https://doi.org/10.1038/238037a0.
3. Chong, M. N., Jin, B., Chow, C. W., & Saint, C. (2010). Recent developments in photocatalytic
water treatment technology: A review. Water Research, 44, 2997–3027.
4. Qian, R., et  al. (2019). Charge carrier trapping, recombination and transfer during TiO2
photocatalysis: An overview. Catalysis Today, 335, 78–90. https://doi.org/10.1016/j.
cattod.2018.10.053.
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 427

5. Dong, S., et  al. (2015b). Recent developments in heterogeneous photocatalytic water treat-
ment using visible light-responsive photocatalysts: A review. RSC Advances, 5, 14610–14630.
https://doi.org/10.1039/C4RA13734E.
6. Mahlambi, M., Ngila, J., & Mamba, B. (2015a). Recent developments in environmental pho-
tocatalytic degradation of organic pollutants: The case of titanium dioxide nanoparticles—A
review. Journal of Nanomaterials, 2015, 790173. https://doi.org/10.1155/2015/790173.
7. Chen, S., & Wang, L.-W. (2012). Thermodynamic oxidation and reduction potentials of photo-
catalytic semiconductors in aqueous solution. Chemistry of Materials, 24, 3659–3666. https://
doi.org/10.1021/cm302533s.
8. Ajmal, A., Majeed, I., Malik, R. N., Idriss, H., & Nadeem, M. A. (2014). Principles and mech-
anisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative over-
view. RSC Advances, 4, 37003–37026. https://doi.org/10.1039/C4RA06658H.
9. Schneider, J., Matsuoka, M., Takeuchi, M., Zhang, J., Horiuchi, Y., Anpo, M., & Bahnemann,
D.  W. (2014). Understanding TiO2 photocatalysis: Mechanisms and materials. Chemical
Reviews, 114, 9919–9986. https://doi.org/10.1021/cr5001892.
10. Walter, M.  G., Warren, E.  L., McKone, J.  R., Boettcher, S.  W., Mi, Q., Santori, E.  A., &
Lewis, N. S. (2010). Solar water splitting cells. Chemical Reviews, 110, 6446–6473. https://
doi.org/10.1021/cr1002326.
11. Yang, M.-Q., Gao, M., Hong, M., & Ho, G.  W. (2018). Visible-to-NIR photon harvesting:
Progressive engineering of catalysts for solar-powered environmental purification and fuel
production. Advanced Materials, 30, 1802894. https://doi.org/10.1002/adma.201802894.
12. Dey, S., Subhasis Patro, S., Suresh Babu, N., Murthy, P. N., & Panda, S. K. (2017). Development
and validation of a stability-indicating RP–HPLC method for estimation of atazanavir sul-
fate in bulk. Journal of Pharmaceutical Analysis, 7, 134–140. https://doi.org/10.1016/j.
jpha.2013.12.002.
13. Yu, J., Kiwi, J., Wang, T., Pulgarin, C., & Rtimi, S. (2019). Duality in the mechanism of hex-
agonal ZnO/CuxO nanowires inducing sulfamethazine degradation under solar or visible light.
Catalysts, 9, 916.
14. Wei, X., Koo, I., Kim, S., & Zhang, X. (2014). Compound identification in GC-MS by simul-
taneously evaluating the mass spectrum and retention index. Analyst, 139, 2507–2514. https://
doi.org/10.1039/c3an02171h.
15. Backes, C.  W., Scheffer, F.  R., Pereira, M.  B., Teixeira, S.  R., & Weibel, D.  E. (2014).
Photosensitised degradation of organic dyes by visible light using riboflavin adsorbed on the
surface of TiO2 nanotubes. Journal of the Brazilian Chemical Society, 25, 2417–2424.
16. Vallejo-Pecharromán, B., Izquierdo-Reina, A., & Luque de Castro, M. D. (1999). Flow injec-
tion determination of chemical oxygen demand in leaching liquid. Analyst, 124, 1261–1264.
https://doi.org/10.1039/A902443C.
17. Matthews, R. W., Abdullah, M., & Low, G. K. C. (1990). Photocatalytic oxidation for total
organic carbon analysis. Analytica Chimica Acta, 233, 171–179. https://doi.org/10.1016/
S0003-­2670(00)83476-­5.
18. Waters, A. (1993). Photocatalysis of TOC measurements. Filtration & Separation, 30,

533–535. https://doi.org/10.1016/0015-­1882(93)80400-­Q.
19. Lin, Y. P., & Mehrvar, M. (2018). Photocatalytic treatment of an actual confectionery waste-
water using ag/TiO2/Fe2O3: Optimization of photocatalytic reactions using surface response
methodology. Catalysts, 8, 409.
20. Zhao, Y., Tao, C., Xiao, G., & Su, H. (2017). Controlled synthesis and wastewater treatment
of Ag2O/TiO2 modified chitosan-based photocatalytic film. RSC Advances, 7, 11211–11221.
https://doi.org/10.1039/C6RA27295A.
21. Parisi, M. L., Fatarella, E., Spinelli, D., Pogni, R., & Basosi, R. (2015). Environmental impact
assessment of an eco-efficient production for coloured textiles. Journal of Cleaner Production,
108, 514–524. https://doi.org/10.1016/j.jclepro.2015.06.032.
22. Forgacs, E., Cserháti, T., & Oros, G. (2004). Removal of synthetic dyes from wastewaters: A
review. Environment International, 30, 953–971. https://doi.org/10.1016/j.envint.2004.02.001.
428 K. R. Davies et al.

23. Ogugbue, C. J., & Sawidis, T. (2011). Bioremediation and detoxification of synthetic wastewa-
ter containing triarylmethane dyes by Aeromonas hydrophila Isolated From Industrial Effluent.
Biotechnology Research International, 2011, 11. https://doi.org/10.4061/2011/967925.
24. Hao, O.  J., Kim, H., & Chiang, P.-C. (2000). Decolorization of wastewater. Critical

Reviews in Environmental Science and Technology, 30, 449–505. https://doi.
org/10.1080/10643380091184237.
25. Lellis, B., Fávaro-Polonio, C. Z., Pamphile, J. A., & Polonio, J. C. (2019). Effects of textile dyes
on health and the environment and bioremediation potential of living organisms. Biotechnology
Research and Innovation, 3, 275–290. https://doi.org/10.1016/j.biori.2019.09.001.
26. Chung, K.-T. (2016). Azo dyes and human health: A review. Journal of Environmental Science
and Health, Part C, 34, 233–261. https://doi.org/10.1080/10590501.2016.1236602.
27. Pereira, L., & Alves, M. (2012). Dyes—Environmental impact and remediation. In

A.  Malik & E.  Grohmann (Eds.), Environmental protection strategies for sustain-
able development (pp.  111–162). Dordrecht: Springer Netherlands. https://doi.
org/10.1007/978-­94-­007-­1591-­2_4.
28. Kant, R. (2012). Textile dyeing industry an environmental hazard. Natural Science., 4(1), 5.
https://doi.org/10.4236/ns.2012.41004.
29. Fujishima, A., Rao, T.  N., & Tryk, D.  A. (2000). Titanium dioxide photocatalysis. Journal
of Photochemistry and Photobiology C: Photochemistry Reviews, 1, 1–21. https://doi.
org/10.1016/S1389-­5567(00)00002-­2.
30. Chatterjee, D., & Dasgupta, S. (2005). Visible light induced photocatalytic degradation of
organic pollutants. Journal of Photochemistry and Photobiology C: Photochemistry Reviews,
6, 186–205. https://doi.org/10.1016/j.jphotochemrev.2005.09.001.
31. Khataee, A. R., & Kasiri, M. B. (2010). Photocatalytic degradation of organic dyes in the pres-
ence of nanostructured titanium dioxide: Influence of the chemical structure of dyes. Journal of
Molecular Catalysis A: Chemical, 328, 8–26. https://doi.org/10.1016/j.molcata.2010.05.023.
32. Sakthivel, S., Shankar, M.  V., Palanichamy, M., Arabindoo, B., & Murugesan, V. (2002).
Photocatalytic decomposition of leather dye: Comparative study of TiO2 supported on alumina
and glass beads. Journal of Photochemistry and Photobiology A: Chemistry, 148, 153–159.
https://doi.org/10.1016/S1010-­6030(02)00085-­0.
33. Anwer, H., Mahmood, A., Lee, J., Kim, K.-H., Park, J.-W., & Yip, A.  C. K. (2019).

Photocatalysts for degradation of dyes in industrial effluents: Opportunities and challenges.
Nano Research, 12, 955–972. https://doi.org/10.1007/s12274-­019-­2287-­0.
34. Qin, N., Wei, W., Huang, C., & Mi, L. (2020). An efficient strategy for the fabrication of
CuS as a highly excellent and recyclable photocatalyst for the degradation of organic dyes.
Catalysts, 10, 40.
35. Chen, X., Wu, Z., Liu, D., & Gao, Z. (2017b). Preparation of ZnO photocatalyst for the effi-
cient and rapid photocatalytic degradation of azo dyes. Nanoscale Research Letters, 12, 143.
https://doi.org/10.1186/s11671-­017-­1904-­4.
36. Saad, A.  M., et  al. (2020). Photocatalytic degradation of malachite green dye using chito-
san supported ZnO and Ce–ZnO nano-flowers under visible light. Journal of Environmental
Management, 258, 110043. https://doi.org/10.1016/j.jenvman.2019.110043.
37. Nguyen, C. H., Fu, C.-C., & Juang, R.-S. (2018). Degradation of methylene blue and methyl
orange by palladium-doped TiO2 photocatalysis for water reuse: Efficiency and degrada-
tion pathways. Journal of Cleaner Production, 202, 413–427. https://doi.org/10.1016/j.
jclepro.2018.08.110.
38. Abdelaal, M.  Y., & Mohamed, R.  M. (2013). Novel Pd/TiO2 nanocomposite prepared by
modified sol–gel method for photocatalytic degradation of methylene blue dye under visible
light irradiation. Journal of Alloys and Compounds, 576, 201–207. https://doi.org/10.1016/j.
jallcom.2013.04.112.
39. Kuo, W. S., & Ho, P. H. (2001). Solar photocatalytic decolorization of methylene blue in water.
Chemosphere, 45, 77–83. https://doi.org/10.1016/S0045-­6535(01)00008-­X.
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 429

40. Reza, K. M., Kurny, A., & Gulshan, F. (2017). Parameters affecting the photocatalytic deg-
radation of dyes using TiO2: A review. Applied Water Science, 7, 1569–1578. https://doi.
org/10.1007/s13201-­015-­0367-­y.
41. Poulios, I., Avranas, A., Rekliti, E., & Zouboulis, A. (2000). Photocatalytic oxidation of
Auramine O in the presence of semiconducting oxides. Journal of Chemical Technology &
Biotechnology, 75, 205–212. https://doi.org/10.1002/(sici)1097-­4660(200003)75:3<205::
aid-­jctb201>3.0.co;2-­l.
42. Wang, N., Li, J., Zhu, L., Dong, Y., & Tang, H. (2008). Highly photocatalytic activity of metal-
lic hydroxide/titanium dioxide nanoparticles prepared via a modified wet precipitation pro-
cess. Journal of Photochemistry and Photobiology A: Chemistry, 198, 282–287. https://doi.
org/10.1016/j.jphotochem.2008.03.021.
43. Ling, C. M., Mohamed, A. R., & Bhatia, S. (2004). Performance of photocatalytic reactors
using immobilized TiO2 film for the degradation of phenol and methylene blue dye present in
water stream. Chemosphere, 57, 547–554. https://doi.org/10.1016/j.chemosphere.2004.07.011.
44. Baran, W., Makowski, A., & Wardas, W. (2008). The effect of UV radiation absorption of cat-
ionic and anionic dye solutions on their photocatalytic degradation in the presence TiO2. Dyes
and Pigments, 76, 226–230. https://doi.org/10.1016/j.dyepig.2006.08.031.
45. Waghmode, T. R., Kurade, M. B., Sapkal, R. T., Bhosale, C. H., Jeon, B.-H., & Govindwar,
S. P. (2019). Sequential photocatalysis and biological treatment for the enhanced degradation
of the persistent azo dye methyl red. Journal of Hazardous Materials, 371, 115–122. https://
doi.org/10.1016/j.jhazmat.2019.03.004.
46. Mahmood, I., Imadi, S. R., Shazadi, K., Gul, A., & Hakeem, K. R. (2016). Effects of pesticides
on environment. In K. R. Hakeem, M. S. Akhtar, & S. N. A. Abdullah (Eds.), Plant, soil and
microbes: Vol. 1. Implications in crop science (pp. 253–269). Cham: Springer International
Publishing. https://doi.org/10.1007/978-­3-­319-­27455-­3_13.
47. Bourguet, D., & Guillemaud, T. (2016). The hidden and external costs of pesticide use. In
E. Lichtfouse (Ed.), Sustainable agriculture reviews. (Vol. 19, pp. 35–120). Cham: Springer
International Publishing. https://doi.org/10.1007/978-­3-­319-­26777-­7_2.
48. Sabarwal, A., Kumar, K., & Singh, R.  P. (2018). Hazardous effects of chemical pesticides
on human health–Cancer and other associated disorders. Environmental Toxicology and
Pharmacology, 63, 103–114. https://doi.org/10.1016/j.etap.2018.08.018.
49. Nicolopoulou-Stamati, P., Maipas, S., Kotampasi, C., Stamatis, P., & Hens, L. (2016).

Chemical pesticides and human health: The urgent need for a new concept in agriculture.
Frontiers in Public Health, 4, 148–148. https://doi.org/10.3389/fpubh.2016.00148.
50. Kanan, S., Moyet, M. A., Arthur, R. B., & Patterson, H. H. (2020). Recent advances on TiO2-­
based photocatalysts toward the degradation of pesticides and major organic pollutants from
water bodies. Catalysis Reviews, 62, 1–65. https://doi.org/10.1080/01614940.2019.1613323.
51. Cruz, M., et al. (2017). Bare TiO2 and graphene oxide TiO2 photocatalysts on the degrada-
tion of selected pesticides and influence of the water matrix. Applied Surface Science, 416,
1013–1021. https://doi.org/10.1016/j.apsusc.2015.09.268.
52. Bhoi, Y. P., Behera, C., Majhi, D., Equeenuddin, S. M., & Mishra, B. G. (2018). Visible light-­
assisted photocatalytic mineralization of diuron pesticide using novel type II CuS/Bi2W2O9
heterojunctions with a hierarchical microspherical structure. New Journal of Chemistry, 42,
281–292. https://doi.org/10.1039/C7NJ03390G.
53. Malato, S., Cáceres, J., Fernández-Alba, A. R., Piedra, L., Hernando, M. D., Agüera, A., &
Vial, J. (2003). Photocatalytic treatment of diuron by solar photocatalysis: Evaluation of main
intermediates and toxicity. Environmental Science & Technology, 37, 2516–2524. https://doi.
org/10.1021/es0261170.
54. Moctezuma, E., Leyva, E., Palestino, G., & de Lasa, H. (2007). Photocatalytic degrada-
tion of methyl parathion: Reaction pathways and intermediate reaction products. Journal
of Photochemistry and Photobiology A: Chemistry, 186, 71–84. https://doi.org/10.1016/j.
jphotochem.2006.07.014.
430 K. R. Davies et al.

55. Ramacharyulu, P.  V. R.  K., Praveen Kumar, J., Prasad, G.  K., & Srivastava, A.  R. (2015).
Synthesis, characterization and photocatalytic activity of Ag–TiO2 nanoparticulate film. RSC
Advances, 5, 1309–1314. https://doi.org/10.1039/C4RA10249E.
56. Chen, H., Shen, M., Chen, R., Dai, K., & Peng, T. (2011). Photocatalytic degradation of com-
mercial methyl parathion in aqueous suspension containing La-doped TiO2 nanoparticles.
Environmental Technology, 32, 1515–1522. https://doi.org/10.1080/09593330.2010.543927.
57. Arévalo Pérez, J. C., et al. (2019). Photocatalytic treatment of pesticides using TiO2 doped
with rare earth. https://doi.org/10.5772/intechopen.84677.
58. Sacco, O., Vaiano, V., Han, C., Sannino, D., & Dionysiou, D. D. (2015). Photocatalytic removal
of atrazine using N-doped TiO2 supported on phosphors. Applied Catalysis B: Environmental,
164, 462–474. https://doi.org/10.1016/j.apcatb.2014.09.062.
59. Belver, C., Han, C., Rodriguez, J. J., & Dionysiou, D. D. (2017). Innovative W-doped titanium
dioxide anchored on clay for photocatalytic removal of atrazine. Catalysis Today, 280, 21–28.
https://doi.org/10.1016/j.cattod.2016.04.029.
60. Miguel, N., Ormad, M. P., Mosteo, R., & Ovelleiro, J. L. (2012). Photocatalytic degradation of
pesticides in natural water: Effect of hydrogen peroxide. International Journal of Photoenergy,
2012, 371714. https://doi.org/10.1155/2012/371714.
61. Vela, N., Calín, M., Yáñez-Gascón, M. J., Garrido, I., Pérez-Lucas, G., Fenoll, J., & Navarro,
S. (2018). Photocatalytic oxidation of six pesticides listed as endocrine disruptor chemicals
from wastewater using two different TiO2 samples at pilot plant scale under sunlight irradia-
tion. Journal of Photochemistry and Photobiology A: Chemistry, 353, 271–278. https://doi.
org/10.1016/j.jphotochem.2017.11.040.
62. Larsson, D. G. J., de Pedro, C., & Paxeus, N. (2007). Effluent from drug manufactures contains
extremely high levels of pharmaceuticals. Journal of Hazardous Materials, 148, 751–755.
https://doi.org/10.1016/j.jhazmat.2007.07.008.
63. Fram, M.  S., & Belitz, K. (2011). Occurrence and concentrations of pharmaceutical com-
pounds in groundwater used for public drinking-water supply in California. Science of The
Total Environment, 409, 3409–3417. https://doi.org/10.1016/j.scitotenv.2011.05.053.
64. Bilgin Simsek, E. (2017). Solvothermal synthesized boron doped TiO2 catalysts: Photocatalytic
degradation of endocrine disrupting compounds and pharmaceuticals under visible light
irradiation. Applied Catalysis B: Environmental, 200, 309–322. https://doi.org/10.1016/j.
apcatb.2016.07.016.
65. Gar Alalm, M., Tawfik, A., & Ookawara, S. (2016). Enhancement of photocatalytic activ-
ity of TiO2 by immobilization on activated carbon for degradation of pharmaceuticals.
Journal of Environmental Chemical Engineering, 4, 1929–1937. https://doi.org/10.1016/j.
jece.2016.03.023.
66. He, Y., Sutton, N. B., Rijnaarts, H. H. H., & Langenhoff, A. A. M. (2016). Degradation of
pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under simulated solar
irradiation. Applied Catalysis B: Environmental, 182, 132–141. https://doi.org/10.1016/j.
apcatb.2015.09.015.
67. Elmolla, E. S., & Chaudhuri, M. (2010). Photocatalytic degradation of amoxicillin, ampicillin
and cloxacillin antibiotics in aqueous solution using UV/TiO2 and UV/H2O2/TiO2 photoca-
talysis. Desalination, 252, 46–52. https://doi.org/10.1016/j.desal.2009.11.003.
68. Jallouli, N., et al. (2018). Heterogeneous photocatalytic degradation of ibuprofen in ultrapure
water, municipal and pharmaceutical industry wastewaters using a TiO2/UV-LED system.
Chemical Engineering Journal, 334, 976–984. https://doi.org/10.1016/j.cej.2017.10.045.
69. Rueda-Marquez, J. J., Levchuk, I., Fernández Ibañez, P., & Sillanpää, M. (2020). A critical
review on application of photocatalysis for toxicity reduction of real wastewaters. Journal of
Cleaner Production, 258, 120694. https://doi.org/10.1016/j.jclepro.2020.120694.
70. Turan, N. B., Erkan, H. S., Engin, G. O., & Bilgili, M. S. (2019). Nanoparticles in the aquatic
environment: Usage, properties, transformation and toxicity—A review. Process Safety and
Environmental Protection, 130, 238–249. https://doi.org/10.1016/j.psep.2019.08.014.
16  Photocatalytic Water Pollutant Treatment: Fundamental, Analysis and Benchmarking 431

71. Blaise, C., Gagné, F., Férard, J.  F., & Eullaffroy, P. (2008). Ecotoxicity of selected nano-­
materials to aquatic organisms. Environmental Toxicology, 23, 591–598. https://doi.
org/10.1002/tox.20402.
72. Chen, G., Feng, J., Wang, W., Yin, Y., & Liu, H. (2017a). Photocatalytic removal of hexavalent
chromium by newly designed and highly reductive TiO2 nanocrystals. Water Research, 108,
383–390. https://doi.org/10.1016/j.watres.2016.11.013.
73. Fostier, A. H., Pereira, M. S. S., Rath, S., & Guimarães, J. R. (2008). Arsenic removal from
water employing heterogeneous photocatalysis with TiO2 immobilized in PET bottles.
Chemosphere, 72, 319–324. https://doi.org/10.1016/j.chemosphere.2008.01.067.
74. Molinari, R., & Argurio, P. (2017). Arsenic removal from water by coupling photocatalysis and
complexation-ultrafiltration processes: A preliminary study. Water Research, 109, 327–336.
https://doi.org/10.1016/j.watres.2016.11.054.
75. Litter, M. I. (2009). Treatment of chromium, mercury, lead, uranium, and arsenic in water by het-
erogeneous photocatalysis. In H. I. de Lasa & B. Serrano Rosales (Eds.), Advances in chemical
engineering (Vol. 36, pp. 37–67). Academic. https://doi.org/10.1016/S0065-­2377(09)00402-­5.
76. Li, Y., Cui, W., Liu, L., Zong, R., Yao, W., Liang, Y., & Zhu, Y. (2016). Removal of Cr(VI) by
3D TiO2-graphene hydrogel via adsorption enriched with photocatalytic reduction. Applied
Catalysis B: Environmental, 199, 412–423. https://doi.org/10.1016/j.apcatb.2016.06.053.
77. Gusain, R., Gupta, K., Joshi, P., & Khatri, O. P. (2019). Adsorptive removal and photocatalytic
degradation of organic pollutants using metal oxides and their composites: A comprehensive
review. Advances in Colloid and Interface Science, 272, 102009. https://doi.org/10.1016/j.
cis.2019.102009.
78. Cheng, L., Liu, S., He, G., & Hu, Y. (2020). The simultaneous removal of heavy metals and
organic contaminants over a Bi2WO6/mesoporous TiO2 nanotube composite photocatalyst.
RSC Advances, 10, 21228–21237. https://doi.org/10.1039/D0RA03430D.
79. Loeb, S. K., et al. (2019). The technology horizon for photocatalytic water treatment: Sunrise
or sunset? Environmental Science & Technology, 53, 2937–2947. https://doi.org/10.1021/acs.
est.8b05041.
80. Dong, S., et  al. (2015a). Recent developments in heterogeneous photocatalytic water treat-
ment using visible light-responsive photocatalysts: A review. RSC Advances, 5, 14610–14630.
https://doi.org/10.1039/C4RA13734E.
Chapter 17
Graphene-Based Photocatalytic Materials:
An Overview

Alex T. Kuvarega, Rengaraj Selvaraj, and Bhekie B. Mamba

17.1  Introduction

Carbon is regarded as the element of life as it can form stable bonds with other car-
bon atoms, oxygen, nitrogen, sulphur and many other elements to form life-­inspired
molecules. Traditionally, only two forms of carbon (graphite and diamond) have
been known [1, 2]. However, recently new synthetic carbon allotropes with out-
standing properties such as carbon nanotubes, carbon nanofibers, fullerenes, carbon
quantum dots and graphene have been discovered. Due to the industrial significance
of most of these carbon allotropes, there has been a lot of interest in pursuing and
capitalising on their desirable properties to fabricate novel materials for different
applications [1, 3, 4].
Among the many carbon allotropes, graphene (GR) has received tremendous
research attention lately due to its exceptional properties and the possibility to
synthesise other modified forms of graphene such as graphene oxide (GO),
exfoliated graphene (EG) and reduced graphene oxide (RGO). Graphene is
composed of network of carbon atoms packed into a honeycomb-like sheet
arrangement [5]. The sheets form stacks on each other to form layered graphene
(Fig. 17.1) [6].
Due to its layered structure, GR is regarded as the basic building block of all-­
dimensional carbon allotropes.
It has found wide applications in the fabrication of photocatalytic heterojunc-
tions for enhanced activity. The interest in the material stems from the enhancement

A. T. Kuvarega · B. B. Mamba
Nanotechnology and Water Sustainability Research Unit, College of Science, Engineering
and Technology, University of South Africa, Florida Science Campus, Florida,
Johannesburg, South Africa
R. Selvaraj (*)
Chemistry Department, Sultan Qaboos University, Muscat, Sultanate of Oman
e-mail: rengaraj@squ.edu.om

© Springer Nature Switzerland AG 2021 433


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_17
434 A. T. Kuvarega et al.

Fig. 17.1  Layers of


graphene sheets. (Adapted
with permission from [6])

Fig. 17.2  Schematic illustration of the graphene as a charge transport medium in the graphene-­
based multicomponent heterojunctions. (Adapted with permission from [7])

in activity due to the formation of Schottky junctions or graphene-cocatalyst hetero-


junctions which improves charge carrier separation through electron transport and
high electrical conductivity of graphene [7]. GR also boosts the surface electrocata-
lytic reactions over cocatalysts due to the enriched electrons in graphene, thus
achieving improved photocatalytic activity (Fig. 17.2). Therefore, any strategy to
enlarge the work function of graphene is beneficial for increasing the Schottky bar-
rier height between graphene and an n-type semiconductor, thus enhancing the pho-
tocatalysis [8].
This chapter provides some highlights on the current research status of GR-based
photocatalysts for various applications.

17.2  Synthesis Methods

A number of methods have been proposed for the synthesis of graphene including
thermal exfoliation, mechanical cleavage, chemical vapour deposition and chemical
exfoliation. The exact method of choice depends on the nature and yield of the
graphene [9–11]. The first attempt at synthesising mono- and multilayer graphite
was reported in 1975 by Lang et al. through thermal decomposition of carbon on Pt
17  Graphene-Based Photocatalytic Materials: An Overview 435

Fig. 17.3  SEM and TEM


morphologies of graphene
sheets. (Adapted with
permission from [15])

substrates [12]. The method was however not aggressively pursued due to
inconsistencies in the quality of the graphene produced. In 2004, Novoselov et al.
were credited for the discovery of graphene in 2004 synthesised through exfoliation
[13, 14]. This technique has gained widespread acceptance, is up to date and remains
the preferred synthesis method of graphene.
In general, the structure and morphology of the resulting GR sheets are the same
irrespective of the synthesis method (Fig. 17.3) [15]. Its low-density fluffy structure
has attracted its application in a wide arrange of areas including water treatment and
environmental decontamination.

17.2.1  Synthesis of Graphene

Exfoliation

Exfoliation of graphite to form graphene through chemical, mechanical or thermal


methods has been reported as one of the easiest and most reliable methods [16, 17].
The method involves intercalation of molecules within the graphite sheets to form a
graphite intercalation compound. This is followed by high-temperature annealing
result in a low-density fluffy material (Fig. 17.4) [18, 19].
Exfoliation has attracted a lot of interest due to potentially easy scale-up to the
industrial production level. However, there is still a need to fine-tune the method in
order to have some control on the number of layers as well as minimising impurity
levels [20].
436 A. T. Kuvarega et al.

Fig. 17.4 (a) SEM image of exfoliated graphite. (b) HRTEM image of few-layered graphene.
(Adapted with permission from [15, 16])

Chemical Vapour Deposition

Chemical vapour deposition (CVD) is a bottom-up solid-phase deposition approach


that involves the annealing of carbon sources on metal catalysts. While this method
is relatively simple, it requires the use of high temperatures to decompose the carbon
sources. The first report on CVD synthesis of graphene involved the pyrolysis of
camphor over a Ni substrate [21]. However, there was still a need to control the
number of layers and minimise the folds. Besides camphor, a mixture of H2 and CH4
has also been used as precursors in CVD and good reproducibility reported [20].

Plasma-Enhanced Chemical Vapour Deposition

In plasma-enhanced chemical vapour deposition (PECVD), a dc plasma is passed


over a Si wafer in the presence of metal substrates (Ni, W, Mo, Zr, Ti, Hf, etc.) and
a gas mixture of CH4 and H2 (0–25% CH4) at a pressure of 10–150 Torr [22, 23].
Different variations of the technique includes the use of microwaves and etching
using atomic hydrogen [23, 24]. Like CVD, PECVD is considered relatively simple
though there is still a need to optimise the technique to control the thickness of the
graphene sheets as well as engineering design for large-scale production [20].

Chemical Methods

Besides the three main methods highlighted above, attempts have also been made to
synthesise graphene using chemical methods. Such methods include liquid-phase
exfoliation where different solvents can be used as graphite intercalants during
exfoliation. Different chemicals have also been used to chemically extract graphene
films from graphite, without the exfoliation step [25]. Sulphuric and nitric acids
17  Graphene-Based Photocatalytic Materials: An Overview 437

have also been intercalated between the layers of graphite followed by rapid
annealing at temperatures up to 1000  °C.  The explosive evaporation of the acid
molecules produces thin graphitic sheets [26]. Many other chemical routes using
different chemical combinations have been reported [27].

Thermal Decomposition of SiC

The thermal decomposition of Si on single crystal of 3H- to 6H-SiC (0001) is


another method that has been reported to produce few-layered graphene sheets [20].
In this method the substrates are subjected to hydrogen etching at 1600 °C for a
short period of time, followed by cooling to below 700  °C.  Upon evacuating
hydrogen from the system, the growth environment is pumped to a pressure of
2 × 10−7 mbar before temperature ramping at a rate of 10–20 °C/min up to a growth
temperature ranging from 1350 to 1650 °C [28–30]. Similarly, other single crystal
substrates such as Ru, Ir, Ni, Co and Pt (0001) have also been reported to be suitable
for the synthesis of graphene by thermal decomposition of ethylene [31].

Unzipping CNTs and Other Methods

Recently a new method of graphene synthesis based on unzipping multiwalled car-


bon nanotubes (MWCNTs) has been reported. In this method, the MWCNTs are
opened up longitudinally through intercalation of lithium and ammonia, followed
by exfoliation in an acid environment and then abrupt heating. A mixture of
nanoribbons, partially opened MWCNTs and graphene flakes are formed [32]. The
unzipping can also be facilitated by plasma etching or multistep chemical treatment
of MWCNTs [33]. The main advantages of this method are simplicity and the
substrate-free capability.

17.2.2  Properties of Graphene

Interest in graphene has been perpetuated by its remarkable and unique properties
such as a high mobility of charge carriers (>200,000 cm2 V−1 s−1), large theoretical
specific surface area (2630 m2g−1), exceptional Young’s modulus values (~1.0 TPa),
large spring constants (1–5  Nm−1), excellent thermal conductivity
(3000–5000 Wm−1 K−1) and optical transmittance (~97.7%) [7, 20, 34]. Graphene is
a 2D semimetal with a zero-gap. It shows a tiny overlap between the valence and the
conduction bands that allows charge carrier movement with little scattering under
ambient conditions. White light absorbance of suspended single-layer graphene is
2.3% with negligible reflectance (~0.1%). Absorbance increases linearly with the
number of graphene layers from 1 to 5. In addition, the dynamic conductivity of
graphene in the visible range [34]. Another very interesting and useful property of
438 A. T. Kuvarega et al.

graphene is the ability to form highly oxidised derivatives with residual epoxides,
hydroxyl and carboxylic acid groups on their surfaces. This has resulted in variants
of graphene such as graphene oxide (GO) and reduced graphene oxide (rGO). The
oxygen-containing reactive surfaces of GO and rGO provide anchor sites for
fabrication of graphene nanocomposites with tunable properties [34].

17.2.3  Preparation of Graphene-Based


Semiconductor Photocatalysts

Recently, numerous studies on the use of GR, GO and rGO as support materials for
photocatalysts have emerged in literature. Many researchers have reported the
coupling of GR, GO and rGO with semiconductors such as TiO2, ZnO, WO3, etc. for
the degradation of organic contaminants [35–37]. The resultant nanocomposites
have found applications in a wide range of fields including environmental
remediation. Various synthetic methods have been proposed for the fabrication of
graphene-based photocatalysts. The most common synthetic routes are the solvo-
thermal, in situ growth, sol-gel and the solution-mixing methods (Fig. 17.5).
In solvothermal method GR or GO is added to the photocatalysts or the photo-
catalyst precursor and the mixture heated in an autoclave. rGO may be formed dur-
ing the solvothermal method. The resulting nanocomposites are washed and dried
prior to application. The in situ method involves mixing the semiconductor precur-
sor with an appropriate GR allotrope to result in nucleation and growth of the semi-
conductor on the GR surface. This is followed by washing, drying and calcination
of the nanocomposite photocatalyst. In the solution-mixing method, GR or GO is
mixed with a photocatalyst precursor under vigorous stirring or ultrasonic agitation
followed by reduction, drying and calcination. The sol-gel method is a common
route to the synthesis of many carbon/semiconductor nanocomposites. In this pro-
cedure, GR or GO is mixed with the metal alkoxide in the presence of an organic
solvent, and the mixture is magnetically stirred to form a sol. The sol is allowed to
age to form a gel which is then dried and calcined. This method results in the forma-
tion of composites with homogeneously dispersed photocatalysts on the surface of
the GR sheets. Other techniques such as CVD, atomic layer deposition (ALD) and
electrochemical deposition have also been used for the preparation of graphene/
semiconductor nanocomposites. These ex situ methods ensure effective chemical
bonding or non-covalent interactions of the GR nanosheets and the pre-­synthesised
photocatalyst nanoparticles [7, 10, 38, 40]. Table 17.1 shows a few examples of the
GR-based photocatalytic nanocomposites and their applications.
17  Graphene-Based Photocatalytic Materials: An Overview 439

Fig. 17.5  Four synthetic strategies for the fabrication of graphene-based photocatalysts. (Adapted
with permission from [38, 39])

17.3  Applications of Graphene-Based Photocatalysts

The world is faced with serious environmental and energy challenges that include
pollution, fossil fuel depletion and global warming. Current efforts to address these
challenges are directed at innovative technologies that can ensure responsible and
sustainable use of available resources. Photocatalysis, a light-driven process of
440 A. T. Kuvarega et al.

Table 17.1  Selected graphene-based photocatalysts


Photocatalyst
Nanocomposite precursor Synthesis method Application References
GO/TiO2 TiO2 nanosheets Microwave and H2-production [41]
hydrothermal
GO/CdS Cd(CH3COO)2 Solvothermal H2 production [42]
GO/BiVO4 Bi(NO3)3 and Hydrothermal Degradation of [43]
NH4VO3 methylene blue
GR/TiO2 TiO2 Hydrothermal Photodegradation of [44]
nanoparticles methylene blue
N-rGO/TiO2 TiO2 Hydrothermal Decomposition of [45]
nanoparticles methyl orange
GR/CdSe Na2SeSO3 and Hydrothermal Degradation of methyl [46]
Cd(CH3COO)2 orange (MO) and
rhodamine B (RhB)
GR/ZnO GR/ Zn(CH3COO)2 Ultrasonication Electrochemical [47]
SnO2 Sn(CH3COO)2 spray pyrolysis applications
GR/WO3 WCl3 Ultrasonication Detection of NO2 [48]
GR/MnO2 KMnO4 Microwave Supercapacitor [49]
synthesis electrodes
GR/Fe2O3 Fe(NO3)3·9H2O Ultrasonication Detection of BPA [50]
N-GR/CdS CdCl2.2.5H2O Ultrasonication H2 production [51]
and Na2S
GR/C3N4 Melamine Impregnation and H2 production [52]
chemical reduction

converting photon energy into chemical energy, has attracted attention during the
past decades as a potentially viable technique for environmental decontamination.
As a green technology, photocatalysis has been widely applied for the degradation
of pollutants, water splitting, reduction of CO2, disinfection and other related
applications. The major limitations of the technique are the low quantum efficiency
and the large band gaps of most photocatalysts. Photocatalysts based on GR and
other photocatalysts have opened up new avenues and pathways for development of
new materials capable of tackling some of the current planetary environmental
challenges.

17.3.1  Photodegradation of Organic Compounds

Graphene has desirable properties that have promoted its use as an electron accep-
tor, transport channel, photosensitizer, supporting material, photocatalyst and cocat-
alyst. GR/photocatalyst nanocomposites have high surface areas that enhance
pollutant adsorption. The composites also exhibit active photocatalytic sites and
band gap narrowing effect [53]. By far the most studied GR/photocatalyst compos-
ites have been used for the photodegradation of pollutants in aqueous environments
17  Graphene-Based Photocatalytic Materials: An Overview 441

and air. Table 17.2 summarises the applications of graphene-based photocatalysts in


photodegradation of different organic compounds.
In these nanocomposite systems, the GR acts as a medium for electron transport,
thereby reducing charge carrier recombination and prolonging the lifetime of the
reactive radicals. GR also absorbs light in the entire UV-Vis region, thus enhancing
visible light absorption in the nanocomposite [68]. A mechanism for the enhancement
of the photoactivity in GR-based systems is presented in Fig. 17.6.

17.3.2  Photocatalytic Water Splitting


and Hydrogen Production

Besides photodegradation of organic pollutants, GR/photocatalyst composites have


also been successfully applied in water splitting to generate hydrogen gas. In these
systems, graphene has been reported to accelerate the charge transfer, surface
reaction kinetics and the thermodynamic properties of the semiconductors.
Cocatalysts such as Pt and Ir have been reported to enhance the photocatalytic water
splitting reaction in GO/Pt/CdS and GO/Ir/TiO2 nanocomposites [42]. The
photoexcitation will result in electrons being transferred from the valence band to

Table 17.2  Photocatalytic activities of some graphene-based photocatalysts


Degradation efficiency
Photocatalyst Irradiation source Pollutant (%) Refs.
GO/ZnO Xenon arc lamp Methyl orange 96.8 [54]
GO/SnO2 500 W tungsten halogen Rhodamine B 99 [55]
lamp
GR/Mn3O4 Visible light >450 nm Methylene 100 [56]
blue
RGO@TiO2 254 nm UV light Methylene 100 [57]
blue
RGO-Ag3PO4 Illumination chamber TOC 97.1 [58]
CuO/RGO 150 W xenon lamp Rhodamine B 100 [59]
GR/Fe3O4 Sunlight 92.43% [60]
GR/NiFe2O4 500 W xenon lamp MB 100% [61]
Ag3PO4/RGO/ 250 W tungsten halogen RhB 98% [62]
Ag lamp
N-GO/Ag2CO3 350 W xenon lamp Phenol 90% [63]
RGO/TiO2 XPA-7 photochemical BPA 58.5% [64]
reactor
GR/BiOBr 300 W tungsten halogen No 40.3% [65]
lamp
RGO/BiVO4 350 W Xe-lamp No 60% [66]
RGO/BiOIO3 150 W tungsten halogen No 50.9% [67]
lamp
442 A. T. Kuvarega et al.

Fig. 17.6  Photocatalytic degradation using graphene TiO2 composite: the addition of shallow trap
and Ti3+ states favours absorption in the visible region. The electron-hole pair generated due to
photoexcitation generates active species like O2− and OH, which degrade the organic pollutants.
(Adapted with permission from [68])

the conduction band of semiconductor, CdS. The electrons will then be channelled


to the cocatalyst (Pt) and some to the graphene sheets where they scavenge H2O and
O2 and any accessible protons to result in the generation of H2 as illustrated in
Fig. 17.7.

17.3.3  Photocatalytic Reduction of CO2

CO2 has been reported as a greenhouse gas responsible for global warming. The
photoreduction and conversion of CO2 into valuable fuels using solar energy are the
innovative strategies to simultaneously solve energy and global warming problems
[69]. Graphene-based semiconductor photocatalysts have also found applications in
this domain. Separation of charge carriers in this process is enhanced through
coupling of multifunctional graphene-based photocatalysts that can work under
visible light irradiation. Formation of the Schottky heterojunction between at least
two materials is reported to enhance CO2 photoreduction due to efficient charge
separation, increased photostability of the semiconductors and enhanced CO2
adsorption [8].
Generally, there are six major advantages of graphene-based photocatalysts for
CO2 reduction:
17  Graphene-Based Photocatalytic Materials: An Overview 443

Fig. 17.7  Schematic illustration of the charge separation and transfer in graphene/CdS system
under visible light. (Adapted with permission from [42])

• Suppressed photogenerated carrier recombination since graphene is a good elec-


tron acceptor and thus under photon irradiation, the photoinduced electrons on
photocatalysts can be channelled to the graphene to induce reduction reactions,
while the holes remain on the photocatalyst for oxidation processes. This results
in enhanced separation of the electron-hole pairs.
• Increased specific surface areas as graphene is known to have an ultra-large theo-
retical specific surface area, and therefore coupling of graphene with a photo-
catalyst can significantly increase the specific surface area of the photocatalyst.
• Increased CO2 adsorption and activation since graphene exhibits a large 2D
π-conjugated structure that is compatible with the CO2 molecules delocalised
π-conjugated system.
• Enhanced photostability of the composite photocatalyst due to the extraordinary
mechanical and chemical stability of graphene.
• Improved nanoparticle dispersion and small nanoparticle sizes as a result of large
volume of surface functional groups which act as anchor sites for the growth of
small-size photocatalysts.
• Enhanced light absorption since graphene can absorb in the entire UV-Vis spec-
trum [70, 71].
In a recent study, highly efficient solvent-exfoliated graphene/P25  TiO2 nano-
composite photocatalyst thin film was synthesised and evaluated for their visible
light photocatalytic reduction of CO2 [72]. In another study, graphene/titania
nanosheet composites with a more intimate interfacial contact showed enhanced
photocatalytic activity for CO2 [73].
444 A. T. Kuvarega et al.

Figure 17.8 illustrates the effect of graphene on enhancing the photocatalytic


reduction of CO2. The photogenerated electron in the conduction band of CuO can
easily transfer to the rGO nanosheets, which inhibits the recombination of electron-­
hole pairs and facilitate the electron transport to the catalytic sites for reduction of
CO2 [74].

17.3.4  Photocatalytic NOx Removal

In recent years, the presence of NOx in the atmosphere has become one of the major
environmental concerns. NOx is responsible for serious environmental problems
such as acid rain, haze formation, photochemical smog, depletion of the ozone layer
and global warming. Besides, NOx is also poisonous and harmful contributing to
adverse human health by causing lung and respiratory problems. For all these
reasons, a plethora of new technologies, including photocatalysis, have been
proposed for the lowering of NOx concentration in the atmosphere [53]. The
presence of oxygen-containing functional groups on the basal plane and sheet edges
of GO that act as anchor sites for covalent and non-covalent interaction with various
molecules has motivated the fabrication of graphene-based photocatalysts for NOx
removal.
In a study on the photocatalytic activity of TiO2/graphene in NOx, oxidation
under UV and visible light irradiation nanocomposite exhibited higher photocatalytic
efficiency than pure TiO2 [75]. In another study, visible light-driven N-doped
(BiO)2CO3/graphene oxide composites also showed improved photocatalytic
activity and selectivity for NOx removal which was ascribed to enhanced oxidation
with •OH and •O2− radicals. N doping played the dominant role of inhibition of NO2
generation [76]. Yang et  al. fabricated activated semi-coke-supported TiO2-rGO
nanocomposite through a one-step solvothermal method for NO removal under

Fig. 17.8  Plausible mechanism of photocatalytic conversion of CO2 into the methanol using rGO-­
CuO nanocomposites under the visible light irradiation. (Adapted with permission from [74])
17  Graphene-Based Photocatalytic Materials: An Overview 445

visible light irradiation. The introduction of rGO was responsible for improved
dispersion, smaller crystalline size, red-shifted absorption band and suppressed
photoinduced charge carrier recombination of TiO2-rGO photocatalysts [77]. In
another recent report, N-doped Bi2O2CO3/graphene quantum dot composite with
enhanced visible light photocatalytic activity for the removal of indoor air pollutant
NO in the ppb level was synthesised. The composite showed high light harvesting
and charge separation efficiency during the photocatalytic process [78].
In a study on the species responsible for the oxidation of NOx, it was reported
that the dominant reactive species of the photocatalytic reaction mechanism are the
photogenerated holes h+ which directly oxidise NO to NO3− (Fig. 17.9) [79].

17.3.5  Photocatalytic Disinfection

Photocatalytic disinfection has received considerable attention as a promising tech-


nology compared to the common disinfection methods such as chlorination, ozone
and UV disinfection due to the in situ generation of strong oxidising radicals, non-
toxicity and photostability [17]. Reactive oxygen species (ROS) such as hydroxyl
radicals (•OH) and superoxide radical anion (•O2−) are reported to inactivate bacte-
ria in water by damaging the cell wall and cell membrane resulting in a leakage of
intracellular components (Fig. 17.10). Coupling graphene with other semiconductor
materials to form heterojunctions results in improved separation of the photogene-
rated charge carriers, increased specific surface area, sufficient active sites for bac-
terial adsorption and increased ROS for bacteria disinfection [80]. In a study, TiO2/
rGO/WO3 nanocomposite synthesised through a hydrolysis-­hydrothermal method
showed enhanced photocatalytic towards inactivating E. coli compared with binary
TiO2/WO3 [81].

Fig. 17.9  Illustration of possible reaction mechanism of the photocatalytic oxidation of NO on the
TiO2/GR. (Adapted with permission from [79])
446 A. T. Kuvarega et al.

Fig. 17.10  Mechanism involved in photo-inactivation of cells by metal oxide/graphene compos-


ite. (Adapted with permission from [81])

Recently, efforts have focussed on visible light-driven photocatalytic bacteria


disinfection through synthesis of solar light-responsive photocatalysts. Graphene
oxide GO/CdS Schottky heterojunctions were recently reported to show effective
anti-microbial activity. Nearly 100% of both Gram-negative E. coli and Gram-­
positive B. subtilis were inactivated within 25 min under visible light irradiation.
The high catalytic performance of the GO/CdS composites was attributed to
effective charge transfer from CdS to GO that reduces the recombination rate of
photogenerated electron-hole pairs, uniform deposition of CdS on GO sheets
which eliminates aggregation of CdS nanoparticles and the strong interactions
between GO and CdS that enhances the durability of the GO/CdS compos-
ites [82].
Besides visible light active materials, nanocomposites comprising of tradition-
ally known anti-microbial metals such as Ag have been incorporated in graphene-
based photocatalysts to enhance the inactivation. The bacterial inactivation
enhancement mechanism in Ag/AgBr/RGO is illustrated in Fig. 17.11. The mecha-
nism consists of two pathways: primary oxidative stress caused by induced reactive
species like •OH and anti-bactericidal effect of released Ag+ ions [83].
17  Graphene-Based Photocatalytic Materials: An Overview 447

Fig. 17.11  Proposed synergistic photocatalytic bacterial inactivation mechanism by plasmonic


Ag/AgBr/0.5% RGO composite photocatalyst. (Adapted with permission from [83])

17.4  Other Applications

17.4.1  Removal of Heavy Metal Ions

The presence of toxic heavy metals even at concentrations below the permissible
limit still causes environmental problems. Therefore, great attention has been paid
to the removal of heavy metal ions from wastewater around the world in recent
years. Photocatalytic reduction has been viewed as a safe, simple, efficient, nontoxic
and economical method to reduce heavy metal ions in aqueous media [84]. Graphene
has also found useful applications in the fabrication of photocatalysts for
photoreduction of heavy metal ions in aqueous media. Zhang et al. reported on the
synthesis and characterisation of TiO2/graphene oxide nanocomposites for
photoreduction of heavy metal ions in reverse osmosis concentrate. High removal
percentages of Cd2+ and Pb2+ were reported. In another study, GO-based materials
showed high conversion rate for removal of toxic Cr(VI) ions in water.
In a study, rGO-decorated raspberry-like TiO2 microspherical composites were
synthesised by a two-step hydrothermal method where TiO2 microspheres were
firstly obtained via hydrothermal reaction and then the TiO2 particles were loaded
on the platform of graphene nanosheet by the second hydrothermal step. The rGO/
TiO2 nanocomposite was evaluated for the photocatalytic reduction of Cr(VI), and
up to 98% efficiency was reported [85]. Guo et  al. reported the remediation of
radioactive wastewater, through the reduction of radioactive hexavalent uranium,
which is commonly found in wastewater from the nuclear industry using a Fe2O3-­
graphene oxide composites. The photocatalytic reduction efficiency of U(VI)
448 A. T. Kuvarega et al.

reached 76.0% after four cycles. The enhanced photocatalytic activity was attributed
to the improved adsorption properties of U(VI) at GO surface and the enhanced
electron transfer from iron oxide to GO [86]. A mechanism for the enhanced activity
was proposed (Fig. 17.12).

17.4.2  Nitrogen Fixation

Ammonia is a very good source of nitrogen for plants and food crops. Currently
nitrogen fixation techniques are based on expensive and environmentally unfriendly
techniques. Nitrogen fixation by semiconductor photocatalysts is an attractive
alternative because of the potential for energy conservation and environment
protection [87]. However, developing highly efficient photocatalysts for N2 fixation
under mild conditions is still a challenge. In photocatalytic nitrogen fixation, the
energy band of semiconductor should allow conditions for reducing N2 and the
conduction band of semiconductor should be more negative than the reduction
potential of the N2 hydrogenation. The use of visible light energy for activating the
photocatalyst will render the technology more energy efficient [88]. Figure 17.13
depicts the possible mechanism for nitrogen fixation on graphene quantum dots
modified flower like Bi2WO6.
Hong Li et al. reported graphene oxide (GO)@polyoxometalate (POM) compos-
ite nanomaterials, with outstanding photocatalytic N2 fixation capability in pure
water. The nanocomposite was reported to facilitate absorption of visible light

Fig. 17.12  Proposed mechanism for the photoreduction of U(VI) in wastewater. (Adapted with
permission from [86])
17  Graphene-Based Photocatalytic Materials: An Overview 449

Fig. 17.13  Proposed mechanism of photocatalytic nitrogen fixation procedure. (Adapted with
permission from [88])

energy to excite abundant photoelectrons to activate N2 reduction, while the rGO


effectively suppressed electron-hole recombination and rapid transfer of electrons
to the absorbed N2 to accelerate NH3 production [89].

17.5  Conclusions and Perspectives

Overall, significant progress has been made in the fabrication of graphene-based


photocatalytic materials through formation of heterojunction with semiconductors.
A wide variety of systems involving different forms of graphene and a wide range
of semiconductor photocatalysts have been studied. The use of UV and visible light
sources to activate the materials has resulted in appreciable success in the various
applications such as degradation of organic pollutants, removal of NOx, photocatalytic
water splitting, carbon dioxide reduction and bacterial inactivation. However, there
is still a need for further studies with regard to the cheap and scalable synthesis of
single-layer or ultrathin graphene materials with much higher surface area. There is
also a need to thoroughly investigate the mechanisms underlying the enhancement
of the photocatalytic activity in the heterojunctions at both the atomic and molecular
levels through advanced characterisation of the structural and electronic properties
of the graphene-based photocatalysts. Evidence of proposed electronic pathways,
charge carrier dynamics and the involvement of the ROS in the photocatalytic
processes need to be backed by verifiable experimental data. Optimisation of the
graphene/photocatalyst ratio in the nanocomposites has not been uniform resulting
in researchers reporting a wide range of different ratios depending on the nature of
their synthesised materials. There is still a need to explore the synthesis of doped
graphene allotropes for enhanced photoactivity. Understanding the effect of specific
dopants on carrier density or band gaps as well as the metallic or semi-conductive
properties of the doped graphene systems is vital for different photocatalytic
450 A. T. Kuvarega et al.

reactions. Nonetheless, there is a glimmer of hope in the development of graphene-­


based photocatalytic materials for environmental photocatalysis. Attention should
now be focussed on the packaging and upscaling of the materials and devices for
real-life applications. Breakthroughs in this field can only be realised through
continued research and development in this fast-growing domain.

Acknowledgements  AT Kuvarega and BB Mamba would like to thank the Nanotechnology and
Water Sustainability Research Unit (NanoWS), the UNISA and the National Research Foundation
(Grants 105891 and 110995) for financially supporting this work.

References

1. Li, X., et al. (2016). Graphene in photocatalysis: A review. Small, 12(48), 6640–6696.
2. Gupta, A., Sakthivel, T., & Seal, S. (2015). Recent development in 2D materials beyond gra-
phene. Progress in Materials Science, 73, 44–126.
3. Choi, W., & Lee, J.-w. (2016). Graphene: Synthesis and applications. CRC Press.
4. Yu, X., et al. (2017). Graphene-based smart materials. Nature Reviews Materials, 2(9), 17046.
5. Rolf, B. (2016). Optical properties of graphene. World Scientific.
6. Nixon, A. Understanding graphene—Part 1. InvestorIntel.
7. Zhang, N., Zhang, Y., & Xu, Y.-J. (2012). Recent progress on graphene-based photocatalysts:
Current status and future perspectives. Nanoscale, 4(19), 5792–5813.
8. Li, X., et al. (2018). Graphene-based heterojunction photocatalysts. Applied Surface Science,
430, 53–107.
9. Zhu, Y., et al. (2010). Graphene and graphene oxide: Synthesis, properties, and applications.
Advanced Materials, 22(35), 3906–3924.
10. Stankovich, S., et al. (2006). Graphene-based composite materials. Nature, 442(7100), 282.
11. Emtsev, K.  V., et  al. (2009). Towards wafer-size graphene layers by atmospheric pressure
graphitization of silicon carbide. Nature Materials, 8(3), 203.
12. Lang, B. (1975). A LEED study of the deposition of carbon on platinum crystal surfaces.
Surface Science, 53(1), 317–329.
13. Novoselov, K., et al. (2005). Two-dimensional atomic crystals. Proceedings of the National
Academy of Sciences, 102(30), 10451–10453.
14. Novoselov, K. S., et al. (2004). Electric field effect in atomically thin carbon films. Science,
306(5696), 666–669.
15. Paronyan, T.  M., et  al. (2017). Incommensurate graphene foam as a high capacity lithium
intercalation anode. Scientific Reports, 7(1), 1–11.
16. Gass, M. H., et al. (2008). Free-standing graphene at atomic resolution. Nature Nanotechnology,
3(11), 676.
17. Xiang, Q., Yu, J., & Jaroniec, M. (2012). Graphene-based semiconductor photocatalysts.
Chemical Society Reviews, 41(2), 782–796.
18. Chung, D. (2016). A review of exfoliated graphite. Journal of Materials Science, 51(1),
554–568.
19. Wang, X., et al. (2009). Large-scale synthesis of few-layered graphene using CVD. Chemical
Vapor Deposition, 15(1–3), 53–56.
20. Choi, W., et al. (2010). Synthesis of graphene and its applications: A review. Critical Reviews
in Solid State and Materials Sciences, 35(1), 52–71.
21. Somani, P. R., Somani, S. P., & Umeno, M. (2006). Planer nano-graphenes from camphor by
CVD. Chemical Physics Letters, 430(1–3), 56–59.
17  Graphene-Based Photocatalytic Materials: An Overview 451

22. Wang, J., et al. (2004). Synthesis of carbon nanosheets by inductively coupled radio-frequency
plasma enhanced chemical vapor deposition. Carbon, 42(14), 2867–2872.
23. Zhu, M., et  al. (2007). A mechanism for carbon nanosheet formation. Carbon, 45(11),

2229–2234.
24. Shang, N. G., et al. (2008). Catalyst-free efficient growth, orientation and biosensing prop-
erties of multilayer graphene nanoflake films with sharp edge planes. Advanced Functional
Materials, 18(21), 3506–3514.
25. Horiuchi, S., et  al. (2004). Single graphene sheet detected in a carbon nanofilm. Applied
Physics Letters, 84(13), 2403–2405.
26. Ndlovu, T., et al. (2014). Exfoliated graphite/titanium dioxide nanocomposites for photodeg-
radation of eosin yellow. Applied Surface Science, 300, 159–164.
27. Park, S., & Ruoff, R. S. (2009). Chemical methods for the production of graphenes. Nature
Nanotechnology, 4(4), 217.
28. Wu, Y., et al. (2008). Top-gated graphene field-effect-transistors formed by decomposition of
SiC. Applied Physics Letters, 92(9), 092102.
29. Iakimova, T., & Yazdi, G.  R. (2018). Fabrication of graphene by thermal decomposition of
SiC. In Epitaxial graphene on silicon carbide (pp. 75–122). Pan Stanford.
30. Mishra, N., et al. (2016). Graphene growth on silicon carbide: A review. Physica Status Solidi
(a), 213(9), 2277–2289.
31. Wintterlin, J., & Bocquet, M.-L. (2009). Graphene on metal surfaces. Surface Science,

603(10–12), 1841–1852.
32. Cano-Marquez, A. G., et al. (2009). Ex-MWNTs: Graphene sheets and ribbons produced by
lithium intercalation and exfoliation of carbon nanotubes. Nano Letters, 9(4), 1527–1533.
33. Kosynkin, D. V., et al. (2009). Longitudinal unzipping of carbon nanotubes to form graphene
nanoribbons. Nature, 458(7240), 872.
34. Huang, X., et  al. (2012). Graphene-based composites. Chemical Society Reviews, 41(2),
666–686.
35. Zhang, H., et al. (2009). P25-graphene composite as a high performance photocatalyst. ACS
Nano, 4(1), 380–386.
36. Fu, D., et  al. (2012). The synthesis and properties of ZnO–graphene nano hybrid for

photodegradation of organic pollutant in water. Materials Chemistry and Physics, 132(2–3),
673–681.
37. Jiang, G., et al. (2011). TiO2 nanoparticles assembled on graphene oxide nanosheets with high
photocatalytic activity for removal of pollutants. Carbon, 49(8), 2693–2701.
38. An, X., & Jimmy, C. Y. (2011). Graphene-based photocatalytic composites. RSC Advances,
1(8), 1426–1434.
39. Radhika, N., et  al. (2019). Recent advances in nano-photocatalysts for organic synthesis.
Arabian Journal of Chemistry, 12, 4550–4578.
40. Bhanvase, B., Shende, T., & Sonawane, S. (2017). A review on graphene–TiO2 and doped gra-
phene–TiO2 nanocomposite photocatalyst for water and wastewater treatment. Environmental
Technology Reviews, 6(1), 1–14.
41. Xiang, Q., Yu, J., & Jaroniec, M. (2011). Enhanced photocatalytic H 2-production activity of
graphene-modified titania nanosheets. Nanoscale, 3(9), 3670–3678.
42. Li, Q., et al. (2011). Highly efficient visible-light-driven photocatalytic hydrogen production
of CdS-cluster-decorated graphene nanosheets. Journal of the American Chemical Society,
133(28), 10878–10884.
43. Gao, L., Qu, F., & Wu, X. (2013). Reduced graphene oxide-BiVO4 composite for enhanced
photoelectrochemical cell and photocatalysis. Science of Advanced Materials, 5(10),
1485–1492.
44. Pan, X., et al. (2012). Comparing graphene-TiO2 nanowire and graphene-TiO2 nanoparticle
composite photocatalysts. ACS Applied Materials & Interfaces, 4(8), 3944–3950.
452 A. T. Kuvarega et al.

45. Xu, Y., et al. (2016). The synergistic effect of graphitic N and pyrrolic N for the enhanced pho-
tocatalytic performance of nitrogen-doped graphene/TiO2 nanocomposites. Applied Catalysis
B: Environmental, 181, 810–817.
46. Ghosh, T., et al. (2013). The characteristic study and sonocatalytic performance of CdSe–gra-
phene as catalyst in the degradation of azo dyes in aqueous solution under dark conditions.
Ultrasonics Sonochemistry, 20(2), 768–776.
47. Lu, T., et al. (2010). Electrochemical behaviors of graphene–ZnO and graphene–SnO2 com-
posite films for supercapacitors. Electrochimica Acta, 55(13), 4170–4173.
48. Srivastava, S., et al. (2012). Faster response of NO2 sensing in graphene–WO3 nanocompos-
ites. Nanotechnology, 23(20), 205501.
49. Yan, J., et al. (2010). Fast and reversible surface redox reaction of graphene–MnO2 composites
as supercapacitor electrodes. Carbon, 48(13), 3825–3833.
50. Wu, C., et  al. (2014). Synergetic signal amplification of graphene-Fe2O3 hybrid and hexa-
decyltrimethylammonium bromide as an ultrasensitive detection platform for bisphenol
A. Electrochimica Acta, 115, 434–439.
51. Jia, L., et al. (2011). Highly durable N-doped graphene/CdS nanocomposites with enhanced
photocatalytic hydrogen evolution from water under visible light irradiation. The Journal of
Physical Chemistry C, 115(23), 11466–11473.
52. Xiang, Q., Yu, J., & Jaroniec, M. (2011). Preparation and enhanced visible-light photocatalytic
H2-production activity of graphene/C3N4 composites. The Journal of Physical Chemistry C,
115(15), 7355–7363.
53. Nikokavoura, A., & Trapalis, C. (2018). Graphene and g-C3N4 based photocatalysts for NOx
removal: A review. Applied Surface Science, 430, 18–52.
54. Yu, M., et al. (2016). Sub-coherent growth of ZnO nanorod arrays on three-dimensional gra-
phene framework as one-bulk high-performance photocatalyst. Applied Surface Science, 390,
266–272.
55. Wei, J., et al. (2016). Synthesis and photocatalytic properties of different SnO2 microspheres
on graphene oxide sheets. Applied Surface Science, 376, 172–179.
56. Madhusudan, P., et  al. (2012). Facile synthesis of novel hierarchical graphene–Bi 2 O 2
CO 3 composites with enhanced photocatalytic performance under visible light. Dalton
Transactions, 41(47), 14345–14353.
57. Cao, Y.-C., et al. (2015). Reduced graphene oxide supported titanium dioxide nanomaterials
for the photocatalysis with long cycling life. Applied Surface Science, 355, 1289–1294.
58. Samal, A., et al. (2016). Reduced graphene oxide–Ag3PO4 heterostructure: A direct Z-scheme
photocatalyst for augmented photoreactivity and stability. Chemistry–An Asian Journal, 11(4),
584–595.
59. Liu, S., et al. (2012). One-pot synthesis of CuO nanoflower-decorated reduced graphene oxide
and its application to photocatalytic degradation of dyes. Catalysis Science & Technology,
2(2), 339–344.
60. Boruah, P.  K., et  al. (2017). Ammonia-modified graphene sheets decorated with magnetic
Fe3O4 nanoparticles for the photocatalytic and photo-Fenton degradation of phenolic com-
pounds under sunlight irradiation. Journal of Hazardous Materials, 325, 90–100.
61. Singh, R., et al. (2017). Nitrogen doped graphene nickel ferrite magnetic photocatalyst for the
visible light degradation of methylene blue. Acta Chimica Slovenica, 64(1), 170–178.
62. Cui, C., et al. (2014). Photo-assisted synthesis of Ag3PO4/reduced graphene oxide/Ag hetero-
structure photocatalyst with enhanced photocatalytic activity and stability under visible light.
Applied Catalysis B: Environmental, 158, 150–160.
63. Song, S., et al. (2017). Construction of Z-scheme Ag2CO3/N-doped graphene photocatalysts
with enhanced visible-light photocatalytic activity by tuning the nitrogen species. Applied
Surface Science, 396, 1368–1374.
64. Luo, L., et al. (2015). Hydrothermal synthesis of fluorinated anatase TiO2/reduced graphene
oxide nanocomposites and their photocatalytic degradation of bisphenol A. Applied Surface
Science, 353, 469–479.
17  Graphene-Based Photocatalytic Materials: An Overview 453

65. Ai, Z., Ho, W., & Lee, S. (2011). Efficient visible light photocatalytic removal of NO
with BiOBr-graphene nanocomposites. The Journal of Physical Chemistry C, 115(51),
25330–25337.
66. Ou, M., et al. (2016). Graphene-decorated 3D BiVO4 photocatalysts with controlled size and
shape for efficient visible-light-induced photocatalytic performance. Materials Letters, 184,
227–231.
67. Xiong, T., et al. (2015). New insights into how RGO influences the photocatalytic performance
of BiOIO3/RGO nanocomposites under visible and UV irradiation. Journal of Colloid and
Interface Science, 447, 16–24.
68. Qiu, B., et al. (2015). Facile synthesis of the Ti 3+ self-doped TiO 2-graphene nanosheet com-
posites with enhanced photocatalysis. Scientific Reports, 5, 8591.
69. Marszewski, M., et  al. (2015). Semiconductor-based photocatalytic CO 2 conversion.

Materials Horizons, 2(3), 261–278.
70. Low, J., Yu, J., & Ho, W. (2015). Graphene-based photocatalysts for CO2 reduction to solar
fuel. The Journal of Physical Chemistry Letters, 6(21), 4244–4251.
71. Ali Tahir, A., et  al. (2016). The application of graphene and its derivatives to energy con-
version, storage, and environmental and biosensing devices. The Chemical Record, 16(3),
1591–1634.
72. Liang, Y. T., et al. (2011). Minimizing graphene defects enhances titania nanocomposite-based
photocatalytic reduction of CO2 for improved solar fuel production. Nano Letters, 11(7),
2865–2870.
73. Liang, Y.  T., et  al. (2012). Effect of dimensionality on the photocatalytic behavior of car-
bon–titania nanosheet composites: Charge transfer at nanomaterial interfaces. The Journal of
Physical Chemistry Letters, 3(13), 1760–1765.
74. Gusain, R., et  al. (2016). Reduced graphene oxide–CuO nanocomposites for photocata-
lytic conversion of CO2 into methanol under visible light irradiation. Applied Catalysis B:
Environmental, 181, 352–362.
75. Trapalis, A., et al. (2016). TiO2/graphene composite photocatalysts for NOx removal: A com-
parison of surfactant-stabilized graphene and reduced graphene oxide. Applied Catalysis B:
Environmental, 180, 637–647.
76. Chen, M., et al. (2018). Visible-light-driven N-(BiO) 2CO3/Graphene oxide composites with
improved photocatalytic activity and selectivity for NOx removal. Applied Surface Science,
430, 137–144.
77. Yang, W., et al. (2015). Solvothermal fabrication of activated semi-coke supported TiO2-rGO
nanocomposite photocatalysts and application for NO removal under visible light. Applied
Surface Science, 353, 307–316.
78. Liu, Y., et al. (2017). N-Doped Bi2O2CO3/graphene quantum dot composite photocatalyst:
Enhanced visible-light photocatalytic no oxidation and in situ drifts studies. The Journal of
Physical Chemistry C, 121(22), 12168–12177.
79. Wang, Y., et  al. (2016). Photocatalytic oxidation of NO over TiO2-Graphene catalyst by
UV/H2O2 process and enhanced mechanism analysis. Journal of Molecular Catalysis A:
Chemical, 423, 339–346.
80. Upadhyay, R. K., Soin, N., & Roy, S. S. (2014). Role of graphene/metal oxide composites as
photocatalysts, adsorbents and disinfectants in water treatment: A review. RSC Advances, 4(8),
3823–3851.
81. Zeng, X., et al. (2017). Highly dispersed TiO2 nanocrystals and WO3 nanorods on reduced
graphene oxide: Z-scheme photocatalysis system for accelerated photocatalytic water disin-
fection. Applied Catalysis B: Environmental, 218, 163–173.
82. Gao, P., et al. (2013). Graphene oxide–CdS composite with high photocatalytic degradation
and disinfection activities under visible light irradiation. Journal of Hazardous Materials, 250,
412–420.
454 A. T. Kuvarega et al.

83. Xia, D., et al. (2016). Synergistic photocatalytic inactivation mechanisms of bacteria by gra-
phene sheets grafted plasmonic AgAgX (X= Cl, Br, I) composite photocatalyst under visible
light irradiation. Water Research, 99, 149–161.
84. Zhang, H., et al. (2018). Synthesis and characterization of TiO 2/graphene oxide nanocom-
posites for photoreduction of heavy metal ions in reverse osmosis concentrate. RSC Advances,
8(60), 34241–34251.
85. Liu, L., et al. (2017). Reduced graphene oxide (rGO) decorated TiO2 microspheres for visible-­
light photocatalytic reduction of Cr (VI). Journal of Alloys and Compounds, 690, 771–776.
86. Guo, Y., et al. (2017). Enhanced photocatalytic reduction activity of uranium (vi) from aque-
ous solution using the Fe 2 O 3–graphene oxide nanocomposite. Dalton Transactions, 46(43),
14762–14770.
87. Fei, T., et al. (2019). Graphene quantum dots modified flower like Bi2WO6 for enhanced pho-
tocatalytic nitrogen fixation. Journal of Colloid and Interface Science, 557, 498–505.
88. Comer, B. M., & Medford, A. J. (2018). Analysis of photocatalytic nitrogen fixation on rutile
TiO2 (110). ACS Sustainable Chemistry & Engineering, 6(4), 4648–4660.
89. Li, X.-H., et  al. (2019). Reduced state of the graphene oxide@ polyoxometalate nanocata-
lyst achieving high-efficiency nitrogen fixation under light driving conditions. ACS Applied
Materials & Interfaces, 11(41), 37927–37938.
Chapter 18
Recent Advances in Nanostructured
Materials for Detoxification of Cr(VI)
to Cr(III) for Environmental Remediation

Udayabhanu, S. B. Patil, and G. Nagaraju

18.1  Introduction

Pollution and climate change are the current most serious issues in the world. The
environment is getting polluted and is in turn affecting the surrounding ecosystem
because of hasty urbanization and industrialization. From the literature, environ-
mental pollution can be denoted as a disagreeable and unwanted amend in chemical,
physical and biological characteristics of water, soil and air that affects living organ-
isms [1]. The pollutants can be either external matters/energies or naturally happen-
ing contaminants which when comes in contact with the environment cause
opposing deviations. There are different types of pollutants, namely, organic, inor-
ganic and biological.
Heavy metals, namely, cobalt, copper, arsenic, selenium, chromium and cad-
mium, are predominantly challenging because they are not decomposable and can
easily be accumulated in living organisms and because of their mutagenic behav-
iour, toxicity and gathering in the human body through the food chain. Even if at
low amount some heavy metals are important micronutrients for animals and plants,

Udayabhanu
Energy Materials Research Laboratory, Department of Chemistry, Siddaganga Institute of
Technology, Tumakuru, Karnataka, India
Center for Research and Innovations, BGSIT, Adichunchanagiri University, B. G. Nagara,
Mandya, Karnataka, India
S. B. Patil
Energy Materials Research Laboratory, Department of Chemistry, Siddaganga Institute of
Technology, Tumakuru, Karnataka, India
Department of Chemistry, The Oxford College of Science, Bengaluru, Karnataka, India
G. Nagaraju (*)
Energy Materials Research Laboratory, Department of Chemistry, Siddaganga Institute of
Technology, Tumakuru, Karnataka, India

© Springer Nature Switzerland AG 2021 455


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_18
456 Udayabhanu et al.

but at higher amount they destructively distress the living organism’s health [2].
These metals can originate from both anthropogenic and natural process and end up
in soil, water, air and their interface.
Among those heavy metals, the Cr metal is widely used in metal alloys, and it is
used in chrome plating. It is extracted by mining process from chromate ore
(FeCr2O4); this is found in different countries such as India, the Philippines,
Kazakhstan, Finland, Zimbabwe and South Africa. In the filed of industries, it has
more value for huge applications because of its high durability; rust resistant coat-
ing and it can be refined to a mirror-like appearance. Cr is unstable; when it is
exposed to air, it instantly forms a chromium oxide layer, and hence it resists form-
ing further oxygen contamination.
Due to the human activities and natural process, the chromium enters into the
environment. The mainstream of chromium III contamination to the environment is
from the textile industries, leather industries and steel industries. In addition to
these, it also enters through the electro painting and chemical manufacturing indus-
trial applications. Contamination of groundwater may be due to the improper dis-
posal of industrial manufacturing equipment and seepage from chromate mines.
Chromium exists in two stable oxidation forms: trivalent Cr(III) and hexavalent
Cr(VI) in natural environment. Cr(VI) is broadly used in the textile dying, anti-­
corrosion, wood preservation, stainless steel, chrome plating and other industries
[3]. When it enters into the water bodies directly from industrial applications it
forms a contamination in water because of improper disposal of mining apparatuses
and manufacturing equipment of industries. Cr(VI) is highly carcinogenic and toxic,
and due to its high solubility in wide range of pH values, it causes health problems
such as pulmonary congestions, liver damage and vomiting, while Cr(III) is non-­
toxic and it can be considered as a human nutrient of lipid and sugar metabolism and
it can be readily precipitated out of the solution in the form of Cr(OH)3 without
needing any complexing agents [4, 5].
When consumed by animals, the effects can include ‘respiratory problems, a
lower ability to fight disease, birth defects, infertility and tumour formation’
(LennTech). This toxic chromium mainly affects the bone, blood and organs and
leads to the mutagenic effect, causing a cancer. Hexavalent chromium is recognized
to be a cancer causing agent only to certain animals, but considered as a carcino-
genic only to the animals of certain circumstances. In general, it is not classified as
carcinogenic according Occupational Safety and Health Administration (OSHA)
and is fairly unregulated, but is considered as level 3 toxicity. Chromium III plays a
vital role to regulate the human vascular and metabolic systems and also in combat-
ing diabetes. Too much chromium III intake may lead to severe skin rash; other
more serious symptoms include allergic reactions, nose irritations/nosebleed, ulcers,
immune system weakening, alteration of genetic material and damage of kidney and
liver, and finally leads to the death of the patient.
However, there is no established limit to human use of chromium III. It has been
documented that individuals had no negative effects if 1000 mg is consumed for a
long time; but, like all the minerals our body needs, excessive consumption can lead
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 457

to poisoning. It is therefore essential to reduce the toxic chromium (VI) to non-toxic


Cr(III).
The WHO approves the enhanced acceptable concentration for the vomit of
Cr(VI) to be 0.1 ppm in surface water and 0.05 ppm into the drinking water. Hence,
the elimination of Cr(VI) from water is essentially required at any probable cost.
Positively, the outstanding evolution in finances and living standard has enhanced
the growth of water and wastewater purification methods [6]. In water bodies, the
most possible Cr(VI) species are CrO42−, Cr2O72−, HCrO4− and H2CrO4, the relative
distribution of which mainly depends on the redox potential, pH of the solution and
concentration of Cr(VI). But, these species do not form insoluble precipitates to
separate by direct precipitation process [7]. The study in India estimated that the
adverse health effects are mainly on the population exposed to high concentrations
of Cr(VI) (∼20 mg/L). Water contaminated with Cr(VI) (maximum levels ranging
between 41 and 156μg/L) showed that a significantly increased incidence leads to
liver cancer mortality (p < 0.001), lung cancer (p = 0.047), and cancers of the kidney
and other genitourinary organs among women (p = 0.025) [8].
To overcome this problem, significant methods were developed for the conver-
sion and detoxification of Cr(VI) ions [9–13]. There are so many nanostructured
materials presently employed for the reduction of Cr(VI) by various methods, which
are discussed below.
The literature survey exposes that the different reductants/adsorbents have been
used to purify the pollutants in aqueous medium. Among them, H2O2, H2S, SO2 and
ferrous ion were stated for chemical reduction of Cr(VI) to Cr(III). But ferrous ions
are not effective in basic medium. Most of the molecules are only degraded under
oxidation process and produce toxic secondary products which may cause environ-
mental extrapolation. H2S and SO2 themselves demonstrate toxicity and generate
extra environmental issues [14]. And also the researchers have observed Cr(VI)
detoxification through biological species like bacteria, yeast, algae, fungi etc., are
relatively moral for Cr(VI) reduction due to less chemical needs and low cost of
operation, but it shows very good performance under laboratory conditions [15].
But these approaches have some notable drawbacks, such as high cost, low effi-
ciency and difficult operations [16]. Hence, a high-performance, low-cost and easy
process for the reduction of chromium present in wastewater is of significant
attention.

18.2  Semiconducting Photocatalysts

Currently, semiconducting natured nanoparticles such as CoFe2O4/ZrO2 [17],


CPVA/MoS2-OH [18], TiO2-modified carbon spheres [19], ZnO-Fe2O3 [20] and
Cu–TiO2/CuO [21] have been continuously followed to release the appearing energy
and environmental disaster, which in concept develops photogenerated species to
perceive solar to chemical energy conversion. Treatment of wastewater through
photocatalysis process is regarded as a new green technology because it consumes
458 Udayabhanu et al.

less energy, there is no secondary pollution and it is easy to operate. Hence,


researches change their consideration to strategy and premise the enhanced photo-
catalysts for completely utilization of the visible light. During photocatalytic Cr
reduction, the light source is irradiated on the surface of the photocatalyst, where
the electrons jump from valence band (VB) to conduction band (CB). The electrons
from the CB are performing the process of reduction of Cr(VI) to Cr(III).

18.2.1  Non-metal Semiconductor-Red Phosphorous

Donghui Li et al. [22] prepared red phosphorus (RP) in nanosize through ultrasonic
and hydrothermal (HU) method and for the first time it is used as photocatalyst for
the reduction of Cr(VI) under visible light irradiation. Nowadays, a class of elemen-
tal semiconductors such as sulphur [23] and silicon [24] has been measured as
promising catalyst under visible light radiations due to their unique advantages of
easy availability, low-price and abundance. When compared to bulk red phosphorus,
nanosized red phosphorus has gained much attention for water splitting and organic
pollutants degradation under visible light irradiation [25–29]. To improve the pho-
tocatalytic activity of red phosphorus, the great efforts have been made such as cata-
lyst dosage [26], composite with other semiconductors [29, 30] and modifying the
surface area and particle size of elemental semiconductors [27, 29]. It confirmed
that the particle size plays a vital part on the photocatalytic activity of red phospho-
rus [31].
Figure 18.1 shows the SEM and TEM images and size distribution of hydrother-
mal and commercial RP. The commercial RP shows a smooth surface morphology
with block-like structure. After hydrothermal treatment, the particle size of red
phosphorus slightly decreased. But hydrothermally synthesized RP shows the rough
surface morphology and reduction in particle size because RP converted into phos-
phoric acid by reacting with water and oxygen [27, 32]. From TEM images, the
particle size of RP from hydrothermally synthesized was found to be 10 nm; it is
comparatively less than that of commercial RP. In ultrasonic process, pressure and
temperature is generated by the sound cavitation; it reduces the particle size of RP,
while ultrasonic initiation is useful to the separation of nano RP [33, 34]. Hence, it
is relatively required method to prepare uniformly distributes and nano-sized RP for
RP-based photocatalyst to degrade water contaminants.
Figure 18.2a displays the graphical abstract of nano-sized RP reducing the
Cr(VI) under visible light and Fig. 18.2b shows that the percentage of Cr(VI) reduc-
tion reached 99% within 15 min in the presence of hydrothermally synthesized RP
(HU-RP) and reached 10% in the presence of commercial RP (C-RP). It revealed
that synthesized RP is more efficient for reduction reaction as compared to C-RP.
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 459

Fig. 18.1  SEM images of (a) commercial RP, (b) hydrothermal RP and (c) hydrothermal ultra-
sonic RP; TEM images of (d) commercial RP and (e) hydrothermal ultrasonic RP; (f) particle size
distribution of hydrothermal ultrasonic RP [22]

18.2.2  Metal Sulphide Semiconductors

In photocatalysis, the very important parameters believed to enhance the catalytic


activity are wide bandgap, suitable reactive sites, more charge separation and trans-
port properties [35]. Tin sulphide (SnS2) is a low-cost, stable material and non-toxic
with a layered structure having a S-SnS triple layer [36]. It has a wide optical band-
gap range from 1.91 to 2.4  eV; hence it acts as a semiconductor photocatalyst
towards the reduction of CO2, biosensing hydrogen generation, photo detectors, etc.
The Nanoflowers and Nanoyarns structure of SnS2 acts as a good photocatalyst in
the reduction of Cr(VI) ions in wastewater and its photocatalytic activity was further
increased by preparing various heterostructures such as SnS2/SnO2 [37]. However,
460 Udayabhanu et al.

Fig. 18.2  Graphical abstract of the photocatalytic Cr reduction by red phosphor under visible light
irradiation (a) and photocatalytic Cr reduction graph (b) [22]

Scheme 18.1  A schematic illustration of the exfoliation of bulk SnS2/n-propylamine into few-­
layer SnS2 nanosheets, hydrogen atoms are omitted for clarity [39]

low efficiency is commonly detected because of the low charge passage and low
separation rate of photogenerated species in SnS2 [38]. Thus, it is serious to advance
method in overcoming such problems by using SnS2 for that application.
Yongping Liu et al. [39] synthesized SnS2 nanosheets through a liquid exfolia-
tion method using SnS2/n-propylamine precursor, as shown in Scheme 18.1. The 2D
structure materials formed via separation of inorganic-organic hybrids which are
made up of single-layered inorganic molecules sandwiched by the single organic
molecule layer through coordinated bonds [40]. Figure  18.3 shows the TEM,
HR-TEM, AFM images of SnS2 nanosheets having the height of about 3.1 nm. The
Cr(VI) reduction reached 96.5% in the presence of SnS2 nanosheets, 48% in the
presence of bulk SnS2 and 6% under SnS2/n-propylamine hybrid under visible light
irradiations. It revealed that SnS2 nanosheets shows more photocatalytic activity
due to its larger surface area.
The process of catalytic reduction of Cr(VI) by SnS2 nanosheets catalyst is simi-
lar to the presence of TiO2 semiconductor catalysts under light source [41], but the
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 461

Fig. 18.3  TEM (a, b), HRTEM (c, d) images, AFM image (e) and the corresponding height profile
(f) for SnS2 nanosheets. The scalebars are 50 nm in (a, b), and 10 nm in (c, d) [39]

percentage of activity is little different. In TiO2, the photoreduction of O2 and Cr(VI)


would ensue instantaneously for the CB of TiO2 (−0.43 V vs. NHE) is more nega-
tive than E(O2/O2−) (−0.33 V vs. NHE) and E(Cr6+/Cr3+) (0.55 V vs. NHE), respec-
tively [42]. While, the CB of SnS2 is −0.32  V, which is between E(O2/O2−) and
E(Cr6+/Cr3+). Thus, under the illumination, the photogenerated electrons on SnS2
surface do not have enough ability to reducing O2, and hence it shared with Cr(VI)
to convert Cr(III). Hence, the schematic representation of reduction of Cr(VI) in the
presence of SnS2nanosheet under light source is projected as follows.
462 Udayabhanu et al.

SnS2 + hv → e − + h + (18.1)

H 2 O + 2 h + → 1 / 2O 2 + 2H + (18.2)

CrO 4 2 − + 8H + + 3e − → Cr 3+ + 4H 2 O (18.3)

The e−-h+ pairs were produced from catalyst under light irradiation (Eq. 18.1).
Then, the e−-h+ travel to the surface of the SnS2 nanosheets in redox reactions. Here,
h+ is involved in the oxidation of H2O (Eq. 18.2) and H+ is produced. At the same
time, e− are involved to reduce Cr(VI) to Cr(III) (Eq. 18.3).
Along with semiconductor nature of nanoparticles, from past year, the carbon
family like graphene has more fascinated due to its nature of excellent electronic,
conductive, high chemical stability, mechanical properties and its large surface area
[43], and also its photocatalytic applications [44]. Nowadays the challenge is to use
these 2D carbon nanomaterials to mimic the narrow energy bandgap semiconduc-
tors such as CdS to form nanocomposites with the functionalities [45]. Rajamathi
et al. [46] and Wang et al. [47] synthesized CdS–graphene hybrid by passing the
H2S gas. Jia et al. [48] synthesized the CdS–N–graphene composites by calcination
and measured that the composite showed a higher catalytic activity in hydrogen
generation under visible light irradiation as compared to bare CdS. Li et al. [49]
demonstrated that synthesis of CdS–graphene composites via solvothermal method
and then acts as high efficient catalyst in the generation H2 gas under visible light
irradiation, it was mainly depends on the presence of graphene which helps as an
electron acceptor and transporter. Even with the above material the photocatalytic
process is not done so far. Hence, move on to the synthesis of CdS-graphene com-
posite material through an inexpensive, quick and versatile technique such as
microwave-­assisted method for enhanced photocatalysis process [50, 51].
Xinjuan Liu et al. [52] synthesized the cadmium sulphide (CdS) and cadmium
sulphide-reduced graphene oxide (CdS-RGO) nanocomposites by microwave-­
assisted reduction method. Figure 18.4 clearly tells the phase and purity of the CdS

Fig. 18.4  XRD patterns of


RGO, CdS and CG-3 [52]
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 463

–2.0

–1.0 e

1.0

h
2.0
V
3.0

Fig. 18.5  Graphical representation showing electron transfer from semiconductor to graphene [52]

and the presence of graphene. Then it was assisted for photocatalytic process in the
reduction of Cr(VI).
Here, the composite CdS-RGO shows an enhanced catalytic activity in the reduc-
tion of Cr(VI); this is due to the incorporation of RGO; it would reduce the recom-
bination of e−/h+ pair and increased the absorption of light intensity [53, 54].
Figure 18.5 shows the graphical representation of the presence of RGO in compos-
ite and stepwise arrangement of energy levels in CdS-RGO composite. The VB of
CdS is 1.55 V and CB is 0.7 V (vs. NHE) and work function of RGO is 0.88 V (vs.
NHE). Due to smaller bandgap, the photo-induced electrons could have transfer
from the CB of CdS to RGO; it helps to separate photo-induced electrons and pro-
mote the charge recombination in e− transfer process.

18.2.3  Metal Sulphide-Metal Sulphide Composite

Xingzhong Yuan et al. [55] prepared the composite of MoS2/Sb2S3 by hydrothermal


method. Figure 18.6a, b displays the TEM and HRTEM images of MoS2 and Sb2S3.
Figure 18.6c, d shows the interplanar spacing 0.53 nm which are corresponding to
the plane (120) of Sb2S3 while 0.62 nm and 0.27 nm are related to the plane of (002)
MoS2 and (100) Sb2S3, respectively [56]. This distinguished interfaces between
MoS2 and Sb2S3 lattice fringes; it revels the formation of MoS2-Sb2S3 heterostruc-
tures; it helps to reduce the recombination of photogenerated species and enhanced
the photocatalytic activity [57]. Then the catalytic activity measured through the
reduction of Cr(VI) under near-infrared radiations and it responds an enhanced pho-
tocatalytic potential compared to bare MoS2 compound.
Here, a novel MoS2/Sb2S3 composite had wide range of energy bandgap and
sulphur vacancies on MoS2; hence it improved the light absorption and also it has
notable photoelectric conversion efficiency [58]. In the composite material, the
band edge potential of MoS2 and Sb2S3 was different as shown in Fig. 18.7. So the
464 Udayabhanu et al.

Fig. 18.6  TEM images of MS-2 (a, b); HRTEM images (c, d) and corresponding elemental map-
ping images (e, f) [55]

Fig. 18.7  Band edge potential of MoS2 and Sb2S3 [55]


18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 465

e− from the CB of MoS2 was easily transferred to CB of Sb2S3, which is utilized for
the reduction of Cr(III).
The Cr(VI) was degraded to 84%, 99% and 72% under UV, visible and NIR light
irradiations in the presence of MoS2/Sb2S3, but bare compounds showed less per-
centage of degradation. In interim, the h+ of the MoS2 is reacted with water to gener-
ate H2O2; it also helps to reduce the Cr(VI) reduction [59, 60].

18.2.4  Metal Oxide Semiconductors

Semiconducting Photocatalyst materials are standing out attributable to its focal


points of being reasonable, stable, and abundantly available in nature. Titanium
dioxide (TiO2) has been generally utilized as a photocatalyst [61]. In any case, TiO2
photocatalyst was not effectively worked under noticeably visible light. To be spe-
cific, the notable impediment of TiO2 is its wide bandgap (3.2 eV) that has reactivity
just under UV illumination [62]. Furthermore, like different photocatalysts (ZnO,
CdS), TiO2 powder type has freshwater ecotoxicity and human danger [63]. To fore-
stall the arrival of TiO2 powder into the earth atmosphere, the creation of a nano-
structure on the glass substrate has developed as an answer [64]. CdS, Ag2S and
PbS/TiO2 were accounted for as noticeable light-initiated photocatalysts to defeat
the constraints of TiO2 [65]. Likewise, in addition, the localized surface plasmon
resonance (LSPR) impact of metal nanoparticles can be utilized to produce photo-
excited electrons. LSPR emerges from the aggregate wavering of free electrons at
the metallic interface or in little metallic nanostructures [66]. Au nanoparticles
(AuNPs) or Ag nanoparticles (AgNPs) were essentially kept on TiO2 to move the
electrons produced from the metal nanoparticles to the semiconductor. These metal-
lic nanoparticles improved the electron move and anticipated the recombination of
holes and electrons, and in this manner, Ag NPs kept on TiO2 indicated commonly
upgraded photocatalytic action. Woocheol Kim et al. [67] synthesized the branched-­
TiO2 microrods on the FTO glass for Photocatalytic reduction of Cr(VI) under
visible-­light irradiation.
The photo-reduction of Cr(VI) to Cr(III) is also carried out in the presence of
AuNPs/B-TiO2/FTO as a catalyst under UV and visible LED light [68]. The reduc-
tion is superior under UV irradiation compared to visible LED light because TiO2
generates more electrons under UV irradiation as shown in Fig. 18.8. By the LSPR
effect more electrons formed which were quickly transferred to the CB of TiO2 [69];
these are involved in the reduction. The reduction process was carried out by adding
formic acid (HCOOH); it acts as both hole scavenger and pH control agent
(Fig.  18.8). As the pH value varies, the reduction mechanism also changes; it is
shown as below.

In acidic medium : Cr2 O7 2 − + 14H + + 6e − → 2Cr 3+ + 7 H 2 O (18.4)



466 Udayabhanu et al.

Fig. 18.8 Schematic
showing the possible
photocatalytic Cr(VI)
reduction mechanism [67]

In neutral medium : CrO 4 2 − + 8H + + 3e − → 2Cr 3+ + 4H 2 O (18.5)



In basic medium : CrO 4 2 − + 4H 2 O + 3e − → 2Cr ( OH )3 + 5OH − (18.6)

Here, the photocatalytic efficiency is more under low-pH medium than high-pH
medium. In basic medium, Cr(OH)3 is deposited on the photocatalyst surface, and
hence decreased its photocatalytic activity [70].
The reduced Cr(III) is adsorbed on the surface of TiO2, because it acts as a posi-
tive charge for the adsorption of Cr2O72− [71]. Figure 18.9 shows the high-resolution
XPS of AuNPs/B-TiO2/FTO catalyst after the reduction of Cr(VI) and adsorption of
Cr(III). These peaks corresponding to the Cr 2p and Ti 2s are observed and were
assigned as Cr(III) and Cr(IV), respectively [72]. It is concluded that Cr(VI) ions
reduced to Cr(III) through photocatalytic process, then Cr(III) ions are deposited on
the surface of AuNPs/B-TiO2/FTO substrates. From this method they can observe
that TiO2 is not detached from AuNPs/B-TiO2/FTO glass in aqueous medium, and it
could not affect extra pollutant, and AuNPs/B-TiO2/FTO was easily removed from
the photocatalytic reaction.
The binding energies for the Cr 2p1/2 and Cr 2p3/2 of AuNPs/B-TiO2/FTO sub-
strate were observed at 587.1 and 577.4 eV, respectively. It is noted that Cr(VI) ions
were reduced to Cr(III) via photocatalytic reduction and then Cr(III) ions were
adsorbed on the AuNPs/B-TiO2/FTO substrates. As results, AuNPs/B-TiO2/FTO
glass was successfully used to visible-light responsive photocatalysts to reduction
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 467

Fig. 18.9  High-resolution XPS of Cr 2p and Ti 2s peaks of the AuNPs/B-TiO2/FTO substrates


after photocatalytic reduction of Cr(IV) to Cr(III) [67]

of Cr(VI) to Cr(III) and removed Cr(III) from aqueous phase. The main environ-
mental importance of this method is TiO2 was not detached from AuNPs/B-TiO2/
FTO glass during aqueous phase reaction; it does not cause additional environmen-
tal pollutants and AuNPs/B-TiO2/FTO substrate was easily recovered after photo-
catalytic reaction.
Atsuhiro Tanaka et al. [73] synthesized Au/TiO2-Pt nanomaterials for oxidation
of water and reduction of Cr(VI) under UV light irradiation. In the photocatalytic
process, during the O2 formation, Ag ions are reduced to Ag metal and these are
deposited on the photocatalyst surface. Hence, the Ag particles defence the light, the
photocatalytic rate of reaction decreases slowly with increasing the reaction time.
This study estimated that the reduction of Cr6+ to Cr3+ along with oxidation of H2O
to O2 is a suitable reaction system to estimate the activity for O2 formation free from
change in photo-absorption of a surface plasmon resonance (SPR)-photocatalyst.
But, Yoneyama et al. have synthesized the TiO2, WO3, Fe2O3 and SrTiO3 and used
for the reduction of Cr6+ at acid medium. Here, the photogenerated e− are involved
for the reduction of Cr6+ to Cr3+ as shown in Eq. 18.7.

Cr2 O7 + 14H + + 6e − → 2Cr 3+ + 7H 2 O (18.7)

The water scavenges like h+ and photogenerated e− were used for the reduction
of Cr6+ ions as shown in Eq. 18.8.

3H 2 O + 6 h + → 3 / 2 O2 + 6H + (18.8)

Final reaction of photocatalysis process is shown below as Eq. 18.9.


468 Udayabhanu et al.

Cr2 O7 2 − + 8H + → 2Cr 3+ + 3 / 2 O2 + 4H 2 O (18.9)

Meanwhile, at pH-8 the Cr3+ is soluble in H2O and so predicted the formation of
Cr is not altered by catalyst and the photocatalytic activity for the oxidation of
3+

H2O in distinction to the reaction system of Ag+-Ag. Here the photocatalytic activity
of Au/TiO2 increased by adding Pt as a co-catalyst in the reduction of Cr6+ to Cr3+
under visible light irradiation.

18.2.5  Metal Oxide-Metal Oxides Composites

Many changes have been researched to enhancing the photoactivity of TiO2 to the
more active visible through doping by Iron, copper and Nitrogen or narrow energy
bandgaped materials like CdSe, Cu2O, Bi2O3 and Fe2O3 [74–79]. WO3 shows out-
standing photochromic, energy storage and electrochromic properties, because the
ionic radius of W6+ is similar to the Ti4+. Hence it could be coupled with TiO2 during
its co-crystallization, succeeding in the formation of composite, i.e. WO3/TiO2 [80–
83]. Here, WO3 could completely hinder the recombination of photogenerated spe-
cies and also decreases the optical energy bandgap of TiO2. Hence, as compared to
bare TiO2, the WO3/TiO2 exhibits enhanced photocatalytic activity [84, 85].
Lixia Yang et al. [86] synthesized the WO3 doped TiO2 nanotube arrays of com-
posites and used as a photocatalyst in the reduction of Cr(VI). Citric acid is used as
a donor scavenger to decrease the photogenerated holes formed by WO3/TiO2 nano-
tubes catalyst, and also it act as a redox agent in increasing the rate of Cr(VI) reduc-
tion. The nanotube composite shows enhanced absorption spectra in visible region
because of its crystalline nature. Hence, it showed highest photoreduction efficiency
in the reduction of Cr(VI) into Cr(III) as compared to TiO2 under solar light
radiations.

18.2.6  Metal Oxide-Graphene

Graphene was discovered in the year of 2004; it has a high electrical conductivity
and large surface area properties. Hence, it has become more wanted in technology
and material science division. Due to its strong pi stacking interactions between the
sheets of graphene decrease its water dispersibility and limit its uses. In the intern-
ing time, Graphene oxide (GO) has more functional groups sited on its edges and
basal planes. Here, those functional groups are detected as well-organized adsor-
bent sites for the elimination of U(VI), Pb(II), Ni(II) and Cr(VI) metal ions. The
separation of GO from aqueous phase is difficult by filtration and centrifugation
methods because of its more hydrophilicity nature and also because its particle size
is too small. Hence, the magnetic nature of Fe2O3 nanoparticles is widely used to
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 469

synthesize the composites with GO, then it acts as a good adsorbent for the elimina-
tion of Cr(VI) and could be easily separated from aqueous phase.
Aiet et al. reported the synthesis of Fe@Fe2O3 core-shell, which were found to
be in nanowires in nano dimensions. 1  g/L of Fe@Fe2O3 core-shell adsorbed
10.06  mg/L of Cr(VI). Here, synthesized nanocomposites show more Cr(VI)
adsorption over the bare components. Wanet al. synthesized the g-Fe2O3@cellulose
aerogel by new template method along with co-precipitation method and used it as
the adsorbent for Cr(VI) removal. Baikousi et al. worked on the synthesis of g-Fe2O3/
carbon hybrids and characterized through different analytical methods and utilized
for the removal of Cr(VI) [87].
Chaopei Kong et al. [87] synthesized the Fe2O3/rGO composites by hydrother-
mal method and demonstrated the removal capacity of Cr(VI) (Scheme 18.2).
The authors have  studied the effect of different temperatures and preparation
time on Cr(VI) reduction ability of Fe2O3/rGO composites. A possible preparation
of Fe2O3/rGO composites and its Cr(VI) reduction mechanism is shown in Scheme
18.3. To check the oxidation states of the different elements of before and after
Fe2O3/rGO used in reduction process was identified by XPS analysis. By this analy-
sis the interactions between the Cr(VI) and the Fe2O3/rGO and GO were identified
as shown in Fig. 18.10.
Figure 18.10a shows the XPS spectrum of the composite, as compared to rGO,
the Fe 2p peaks of the Fe2O3/rGO (before reduction) seemed at the 711.4 and
724.9 eV (inserted Fig. 18.6a). The doublet Fe 2p peak of Fe2O3 with 13.5 eV split-
ting energy, indicating that Fe2O3 is positively attached on GO and also confirmed
the attach through satellite peak situated at 719.5 eV. The peaks at 577.5 eV and
587.2 eV consistent to Cr 2p3/2 and Cr2p1/2, which is detected in Fig. 18.10b.
These two peaks corresponding to Cr(III) and Cr(VI), it indicated that the mod-
erate reduction of Cr(VI) to Cr(III) was carried out. Figure 18.10c, d shows the C1s
spectra of before and after adsorption of Fe2O3/rGO, respectively. The C functional
groups of before and after adsorption of Fe2O3/rGO have been decomposed into
C–C, C–H, C–OH and C-OOH. In Fig. 18.10d the percentage of C-OH and C-OOH
decreased, signifying that these groups are intricate in the reaction, and also similar
result has been found from Fig. 18.10e, f.

Scheme 18.2  Schematic illustration for the preparation of Fe2O3/rGO and removal mecha-
nism [87]
470 Udayabhanu et al.

Scheme 18.3 (a) Schematic diagram for the synthesis of BT-COFs; (b) top and (c) side views of
TPB-BT-COF; (d) top and (e) side views of TAPT-BT-COF; (f) photographs of TPB-BY-COF,
TAPT-BT-COF and TPB-TP-COF [98]

The Fe2O3/rGO have the nature of attracting Cr(VI) anions via electrostatic
forces with carboxyl and protonated hydroxyl groups as shown in Eqs.  18.10
and 18.11.

SurfaceOH 2 + + HCrO 4 − ↔ surface OH 2 CrO4 − + H + (18.10)

Surface COOH 2 + + HCrO 4 − ↔ surfaceCOOH 2 CrO 4 − + H + (18.11)


HCrO 4 − + 7H + + 3e − ↔ Cr 3+ + 4H 2 O (18.12)

Surface FeOH + HCrO 4 − ↔ surface FeCrO 4 − + H 2 O (18.13)



Then the Cr(VI) get reduced to Cr(III) by hydroxyl groups, which acts as an
electron donor; it is observed in Eq. 18.11. Some part of reduced Cr(III) is entered
into aqueous solution, the remaining part get precipitated on the active site of the
Fe2O3/rGO. The Fe2O3 nanoparticles were vastly distributed due to larger surface of
rGO. Hence, it is concluded that Fe2O3/rGO contain Fe-OH groups; these help in the
removal of Cr(VI), which is shown in Eqs. 18.11–18.13.

18.2.7  Molecular Organic Frameworks (MOFs)

The metal sulphides and metal oxides suffer from the quick recombination of e−/h+
pair, easy agglomeration and the low efficiency of utilization of solar energy. Hence,
developing a material showing enhancing photocatalytic activity is desirable. Due
to interesting properties of MOFs and coordination polymers, scientists have sug-
gested significantly for consideration in material science field [88, 89]. Recently,
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 471

Fig. 18.10  The spectra survey scans of GO, Fe2O3/rGO and used Fe2O3/rGO (a); Cr 2p spectra of
used Fe2O3/rGO (b); C 1s spectra of unused Fe2O3/rGO (c); C 1s spectra of used Fe2O3/rGO (d); O
1s spectra of unused Fe2O3/rGO (e); O 1s spectra of used Fe2O3/rGO (f) [87]

some researchers have reported on the synthesis and photocatalytic activity of


MOFs materials through the reduction of Cr(VI) and degradation of organic dyes by
utilizing its photo-induced electrons. Zhao et al. [85] have revealed the AOPs for
Rhodamine B organic dye in the presence of MOFs; here highly reactive oxidants,
i.e. ozone and hydroxyl radicals, are involved in the degradation of organic dye
[90–94]. Wang et  al. showed the reduction of Cr(VI) by NH2-MIL-125(Ti); it
showed high efficiency of photocatalytic activity [95]. And, MIL-68(In)-NH2 MOFs
472 Udayabhanu et al.

also have been established to be attractive catalyst in the reduction of Cr(VI) under
light source [96].
These results prove that the MOFs could act as a potential photocatalyst in reduc-
tion of chromium and in the degradation of organic dyes, but some MOFs are unsuc-
cessful in carrying out the particular photocatalytic reactions, mainly due to the
deficiency of absorbance sites in the MOFs compounds.
Hongmei Zhao et al. [97] worked on the synthesis of novel MOFs via a control-
lable pillared-layer method using bipyridine as a pillared ligand. The experimental
analysis showed that pillared-layer MOF acts as an efficient catalyst under visible
light in the reaction of Cr(VI) reduction and organic dye degradation (Fig. 18.11).
After the addition of oxidant and hole scavenger in MOFs approached catalytic
reaction, the enhanced photocatalytic efficiency is observed.
The two-dimensional (2D)covalent organic frameworks (COFs) assist as a dif-
ferent stage for improving novel photocatalyst and conductive materials because of
its well arrangement of conjugated structure and separated displays of columns
[92, 93].
Addition of benzothiadiazole (BT) (strong electron deficient) molecule build a
two highly porous and crystalline BT-COFs via Schiff base (acid-catalysed) reac-
tion as shown in the below Scheme 18.3. Electron deficient nature of BT could
basically tune the optical energy bandgaps and increase the charge separation prop-
erty of conjugated polymers. The photocatalytic reduction of Cr(VI) by the COFs is
explained in Fig. 18.12 [98].

Fig. 18.11 (a) Pillared-layer structure; (b) layer structure constituted by BPDC ligand and zinc
dimer; (c) a view shows the pillaring ligand coordinated to layer in the structure. Blue, red, purple
and orange spheres/polyhedral represent nitrogen, oxygen, zinc and carbon atoms, respectively.
All hydrogen atom and guest molecules are omitted for clarity [97]
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 473

Fig. 18.12  The probable mechanism of photoreduction of Cr(VI) by BT-COFs [98]

18.2.8  Cr(VI) Reduction by Bacterias

Chromium containing industrial wastewater is currently treated by ion-exchange


and chemical reduction methods, due to the more cost for the ion exchange resin,
but chemical method is still in use for disposing more concentrated Cr(VI) contain-
ing water. As of late, the utilization of a natural strategy to remediate chromate
infected wastes has been viewed as a promising, safe and financially savvy innova-
tion, particularly when managing Cr(VI) containing wastewater at low-to-mid fixa-
tions (10–200  mg/L). Different organisms were found to be able to diminish
profoundly lethal Cr(VI) to less dangerous Cr(III) [99]. Some previous specialists
accepted that such movement can ‘fix’ Cr(VI) into an insoluble structure, for exam-
ple, Cr(OH)3, that can only with significant effort be scattered. In any case, late
examinations uncovered that the microbial reduction of Cr(VI) may be soluble in
organoCr(III) building blocks. In oxidative conditions, the versatile organoCr(III)
buildings can be reoxidized into Cr(VI).This is a pertinent deterrent to actualize
industrial utilizations of chromate reducing microscopic organisms. Thus, how to
immobilize the reduced Cr(III) in microbial frameworks turns into a test and simul-
taneously an important research subject. Yangjian Cheng et al. [99] estimated that it
is the particular coordination of Cr(III) to the dissolvable organic molecules in the
bacterial culture medium that represses compelling immobilization of Cr(III) on the
cells accomplished effective immobilization of Cr(VI) as Cr(III) by O. anthropi and
Planococcus citrus in 5–50 L pilot-scale tests.
474 Udayabhanu et al.

Dong-Hun Kim et  al. [100] contemplated the accumulation of amorphous


Cr(III)-Te(IV) nanoparticles on the surface of Shewanella oneidensis MR-1 through
reduction of Cr(VI). They have the outcomes that the presence of both harmful
Cr(VI) and Te(IV), S. oneidensis MR-1 was reduced Cr(VI) to the less poisonous
Cr(III) structure however not ready to reduce Te(IV) to Te(0). Subsequently, the
decreased Cr(III) complexed with Te(IV) and encouraged as undefined Cr(III)-
Te(IV) nanoparticles on the bacterial cell surface.
These outcomes may give another way to deal with co-removal of poisonous
Cr(VI) and Te(IV) as steady stable solid Cr(III)-Te(IV) buildings through ecologi-
cally well-disposed procedures. Electron microscopy investigations were performed
to decide morphology and the limitation of the accelerated Cr and Te in the simul-
taneous culture containing both Cr(VI) and Te(IV) with S. oneidensis MR-1. SEM
pictures demonstrated that the bacterial cell surface was thickly secured with
nanoparticles at 24 h hatching periods (Fig. 18.13a). The nanoparticles were equita-
bly scattered on the cell surface (Fig. 18.13b). Interestingly, the cell surface stayed
clear after a similar hatching time upon treatment with Cr(VI) alone. Totals were
showed up around cells in these societies after hatching up to 120 h. In past reports,
the areas of reduced Cr(III) accumulation in S. oneidensis MR-1 societies were dif-
fered relying upon test conditions [101–103]. S. oneidensis MR-1 has been accounted
and equipped for lessening lethal Cr(VI) to less harmful insoluble Cr(III) nanopar-
ticles on the bacterial cell surface by c-type cytochromes, MtrC and OmcA [101,
104]. Besides, little organic framework or useful functional groups on microbial cell
surfaces have been proposed to assume a significant job in metal authoritative,
which may affect the development of dissolvable Cr(III) final results [105]. As
recently detailed, cell cultures containing Te(IV) alone showed intracellular assem-
bly of Te (0) nanorods [106].
Hai-Kun Zhang et al. [107] examined the Cr(VI) reduction and Cr(III) immobi-
lization by Acinetobacter sp. HK-1 with the assistance of a novel quinone/graphene
oxide composite. They have arranged the materials of 2-aminoanthraquinone-GO
(AQ-GO) and 2-amino-3-chloro-1,4-naphthoquinone-GO (NQ-GO). The genuine
component engaged with this procedure is, under anaerobic conditions, the general
Cr(VI) bioreduction and Cr(III) precipitation at pH 7.5 and are shown in Eqs. 18.14
and 18.15, individually. At the point when glucose was utilized as an electron con-
tributor (Fig. 18.14), the Cr(VI) bioreduction is depicted in Eq. 18.16, which was
accounted by Chen et al. [108].

CrO 4 2 − (aq ) + 8H + (aq ) + 3e − → Cr 3+ (aq ) + 4H 2 O (18.14)



Cr 3+ (aq ) + 4H 2 O → Cr ( OH )3(S) + 3H + (aq ) + H 2 O (18.15)

C6 H12 O6 + 8CrO 4 2 − (aq ) + 14H 2 O → 8Cr ( OH )3(s) + 10OH − + 6HCO− (aq ) (18.16)

As appeared in Eq. 18.16, Cr(III) was hypothetically immobilized into the type
of insoluble Cr(OH)3; remaining Cr(III) in the liquid mixture was as yet released
with the effluents [109, 110].
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 475

Fig. 18.13  SEM (a), magnified SEM (b), TEM (c), magnified TEM (d), thin section TEM (e)
images with SAED patterns (d inserted), and EDS spectra (f) of nanoparticles formed in concur-
rent cultures containing both Cr(VI) and Te(VI) and Te (IV) in the presence of S. oneidensis MR-1
at 24 h of incubation [100]

Truth be told, when natural substances existed in medium or were produced by


microscopic organisms, exceptionally soluble organo-Cr(III), finished results would
be delivered [111]. This organo-Cr(III) items are moderately general and stable dur-
ing Cr(VI) bioremediation. Past examinations demonstrated that bacterial plank-
tonic cells could just immobilize a piece of the framed Cr(III) during Cr(VI)
bioreduction. The biofilms are progressively fit for Cr(III) immobilization con-
trasted and planktonic cells, while Cr(VI) drop rate by biofilms is lower than that by
planktonic cells [112]. In the present investigation, no Cr(III) was identified in the
supernatant of the solution mixture. Additionally, the designed Cr(III) concentration
in pellets was like the concentration of initial Cr(VI) (Fig.  18.15). This outcome
shows that strain HK-1 can successfully immobilize Cr(III) utilizing planktonic
476 Udayabhanu et al.

Fig. 18.14  Schematic representation for the biological chromium reduction and their images after
11 h of treatment [107]

cells. From a down-to-earth perspective, the insoluble product results are attractive
in light of the fact that they can be effectively removed from reaction mixtures.
Based on the above examination, the tool of Cr(VI) reduction and Cr(III) immobili-
zation by strain HK-1 with the help of NQ-GO was proposed: (I) Direct reduction.
Cr(VI) was reduced to Cr(III) straightforwardly and insoluble Cr(III) was immobi-
lized by glycolipids emitted by strain HK-1; (II) Indirect decrease.
To start with, NQ-GO was diminished by external film proteins of strain HK-1;
second, the shaped hydroquinones reduced Cr(VI) in a simply chemical redox
response; and last, insoluble Cr(III) was immobilized by glycolipids discharged by
strain HK-1.
The outcomes demonstrated that the surfaces of the Cr(VI) treated cells were
harsh and that the chromium component was distinguished on the cell surfaces
(Fig.  18.15a, c). For Cr(VI) untreated cells, their surfaces were smooth, and no
chromium was seen on the cell surfaces (Fig. 18.15b, d). The greyish-green precipi-
tation was additionally demonstrated to contain Cr(III) utilizing XPS in the NQ-GO
material (Fig. 18.15e). In the XPS spectrum, a couple of peaks at 574.9 and 584.4 V
speaking to Cr3+ 2p3/2 and Cr3+ 2p1/2, respectively, were watched, which was steady
with the examination. The SEM investigation proposes that strain HK-1 may emit
some clingy metabolites situated on its cell surface within the sight of Cr(VI)
(Fig. 18.15a).
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 477

(a) (b)

10 μm 10 μm

(c) (d)
C
O
Cr
C
Ca
Ca
N
P O
Ca Cr P
Cr Fe

0 2 4 6 8 kev 0 2 4 6 8 kev
3+
NK-GO Cr 2p3/2
4000 3+
Cr 2p1/2

3600
Intersity (a.u.)

3200
Abs

Control
2800
Cells
2400
(e) (f) Cells+Cr(VI)
Cells+Cr(VI)+NQ-GO
590 585 580 575 570 565 4000 3500 3000 2500 2000 1500 1000 500
Binding Energy (eV) Wavelength (cm-1)

Fig. 18.15  Characterization of the reduced products. (a) and (c) are the SEM-EDX analysis of the
reduced products in NQ-GO-free system; (b) and (d) are the SEM-EDX analysis of the reduced
products in the NQ-GO supplemented system; (e) XPS analysis of the reduced products; and (f)
FTIR analysis of the reduced products [107]

18.3  Conclusion

In this chapter, we have discussed in detail the physical and chemical properties of
the chromium along with the effect on plant and animal kingdoms. Furthermore, we
have discussed the reduction of toxic Cr(VI) to Cr(III) by the different mechanism
pathways efficiently by chemical and biological methods. All these methods are
effectively employed for the reduction of Cr(VI) by their own way. Chemical meth-
ods are quite faster, economical and more efficient compared to biological methods.
478 Udayabhanu et al.

References

1. Ko, L. Y., Qin, Y. Y., & Wong, M. H. (2012). Heavy metal overloads and autism in children
from mainland China and Hong Kong: A preliminary study. In Environmental contamination-­
health risks and ecological restoration (pp. 35–46).
2. Barrera-Díaz, C. E., Lugo-Lugo, V., & Bilyeu, B. (2012). A review of chemical, electrochem-
ical and biological methods for aqueous Cr (VI) reduction. Journal of Hazardous Materials,
223, 1–12.
3. Wojtyła, S., & Baran, T. (2018). Insight on doped ZnS and its activity towards photocata-
lytic removing of Cr (VI) from wastewater in the presence of organic pollutants. Materials
Chemistry and Physics, 212, 103–112.
4. Eary, L., & Rai, D. (1988). Chromate removal from aqueous wastes by reduction with ferrous
ion. Environmental Science & Technology, 22(8), 972–977.
5. Yurik, T., & Pikaev, A. (1999). Radiolysis of weakly acidic and neutral aqueous solutions of
hexavalent chromium ions. High Energy Chemistry, 33(4), 208–212.
6. Wang, C.-C., Du, X.-D., Li, J., Guo, X.-X., Wang, P., & Zhang, J. (2016). Photocatalytic
Cr (VI) reduction in metal-organic frameworks: A mini-review. Applied Catalysis B:
Environmental, 193, 198–216.
7. Nriagu, J. O., & Nieboer, E. (1988). Chromium in the natural and human environments. Wiley.
8. Sun, H., Brocato, J., & Costa, M. (2015). Oral chromium exposure and toxicity. Current
Environmental Health Reports, 2(3), 295–303.
9. Shevchenko, N., Zaitsev, V., & Walcarius, A. (2008). Bifunctionalized mesoporous silicas
for Cr (VI) reduction and concomitant Cr (III) immobilization. Environmental Science &
Technology, 42(18), 6922–6928.
10. Zhang, Y., Zhang, D., Zhou, L., Zhao, Y., Chen, J., Chen, Z., & Wang, F. (2018). Polypyrrole/
reduced graphene oxide aerogel particle electrodes for high-efficiency electro-catalytic syn-
ergistic removal of Cr (VI) and bisphenol A. Chemical Engineering Journal, 336, 690–700.
11. Daṃbrowski, A., Hubicki, Z., Podkościelny, P., & Robens, E. (2004). Selective removal
of the heavy metal ions from waters and industrial wastewaters by ion-exchange method.
Chemosphere, 56(2), 91–106.
12. Li, C., Yu, J., Li, W., He, Y., Qiu, Y., Li, P., Wang, C., Huang, F., Wang, D., & Gao, S. (2018).
Immobilization, enrichment and recycling of Cr (VI) from wastewater using a red mud/car-
bon material to produce the valuable chromite (FeCr2O4). Chemical Engineering Journal,
350, 1103–1113.
13. Liu, Q., Liu, Q., Liu, B., Hu, T., Liu, W., & Yao, J. (2018). Green synthesis of tannin-­
hexamethylendiamine based adsorbents for efficient removal of Cr (VI). Journal of
Hazardous Materials, 352, 27–35.
14. Kendelewicz, T., Liu, P., Doyle, C., & Brown, G., Jr. (2000). Spectroscopic study of the reac-
tion of aqueous Cr (VI) with Fe3O4 (111) surfaces. Surface Science, 469(2–3), 144–163.
15. Srivastava, N., & Majumder, C. (2008). Novel biofiltration methods for the treatment of
heavy metals from industrial wastewater. Journal of Hazardous Materials, 151(1), 1–8.
16. Zhang, D., Li, X., Tan, H., Zhang, G., Zhao, Z., Shi, H., Zhang, L., Yu, W., & Sun, Z. (2014).
Photocatalytic reduction of Cr (VI) by polyoxometalates/TiO 2 electrospun nanofiber com-
posites. RSC Advances, 4(84), 44322–44326.
17. Emadian, S. S., Ghorbani, M., & Bakeri, G. J. S. M. (2020). Magnetically separable CoFe2O4/
ZrO2 nanocomposite for the photocatalytic reduction of hexavalent chromium under visible
light irradiation. Synthetic Metals, 267, 116470.
18. Wang, K., Chen, P., Nie, W., Xu, Y., & Zhou, Y. (2019). Improved photocatalytic reduction
of Cr (VI) by molybdenum disulfide modified with conjugated polyvinyl alcohol. Chemical
Engineering Journal, 359, 1205–1214.
19. Swain, S., & Basu, S. (2020). Visible light-driven photocatalytic reduction of chromium (VI)
using TiO 2-modified carbon sphere hybrid photocatalyst. Energy, Ecology and Environment,
6.1(2021), 26–34.
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 479

20. Dhiman, P., Sharma, S., Kumar, A., Shekh, M., Sharma, G., & Naushad, M. J. C. I. (2020).
Rapid visible and solar photocatalytic Cr (VI) reduction and electrochemical sensing of dopa-
mine using solution combustion synthesized ZnO–Fe2O3 nano heterojunctions: Mechanism
elucidation. Ceramics International, 46(8), 12255–12268.
21. Reddy, N. L., Shankar, M., Sharma, S., & Nagaraju, G. (2020). One-pot synthesis of Cu–
TiO2/CuO nanocomposite: Application to photocatalysis for enhanced H2 production, dye
degradation & detoxification of Cr (VI). International Journal of Hydrogen Energy, 45(13),
7813–7828.
22. Li, D., Li, J., Jin, Q., Ren, Z., Sun, Y., Zhang, R., Zhai, Y., & Liu, Y. (2019). Photocatalytic
reduction of Cr (VI) on nano-sized red phosphorus under visible light irradiation. Journal of
Colloid and Interface Science, 537, 256–261.
23. Liu, G., Niu, P., Yin, L., & Cheng, H.-M. (2012). α-Sulfur crystals as a visible-light-active
photocatalyst. Journal of the American Chemical Society, 134(22), 9070–9073.
24. Fellahi, O., Barras, A., Pan, G.-H., Coffinier, Y., Hadjersi, T., Maamache, M., Szunerits, S., &
Boukherroub, R. (2016). Reduction of Cr (VI) to Cr (III) using silicon nanowire arrays under
visible light irradiation. Journal of Hazardous Materials, 304, 441–447.
25. Yuan, Y.-P., Cao, S.-W., Liao, Y.-S., Yin, L.-S., & Xue, C. (2013). Red phosphor/g-C3N4
heterojunction with enhanced photocatalytic activities for solar fuels production. Applied
Catalysis B: Environmental, 140, 164–168.
26. Dang, H., Dong, X., Dong, Y., Fan, H., & Qiu, Y. (2014). Enhancing the photocatalytic H 2
evolution activity of red phosphorous by using noble-metal-free Ni (OH) 2 under photoexci-
tation up to 700 nm. RSC Advances, 4(84), 44823–44826.
27. Li, W., Yue, J., Hua, F., Feng, C., Bu, Y., & Chen, Z. (2015). Enhanced visible light photo-
catalytic property of red phosphorus via surface roughening. Materials Research Bulletin,
70, 13–19.
28. Xia, D., Shen, Z., Huang, G., Wang, W., Yu, J. C., & Wong, P. K. (2015). Red phosphorus: An
earth-abundant elemental photocatalyst for “green” bacterial inactivation under visible light.
Environmental Science & Technology, 49(10), 6264–6273.
29. Hu, Z., Yuan, L., Liu, Z., Shen, Z., & Yu, J. C. (2016). An elemental phosphorus photocatalyst
with a record high hydrogen evolution efficiency. Angewandte Chemie International Edition,
55(33), 9580–9585.
30. Shi, Z., Dong, X., & Dang, H. (2016). Facile fabrication of novel red phosphorus-CdS com-
posite photocatalysts for H2 evolution under visible light irradiation. International Journal of
Hydrogen Energy, 41(14), 5908–5915.
31. Ansari, S. A., Ansari, M. S., & Cho, M. H. (2016). Metal free earth abundant elemental red
phosphorus: A new class of visible light photocatalyst and photoelectrode materials. Physical
Chemistry Chemical Physics, 18(5), 3921–3928.
32. Wang, F., Li, C., Li, Y., & Jimmy, C. Y. (2012). Hierarchical P/YPO4 microsphere for pho-
tocatalytic hydrogen production from water under visible light irradiation. Applied Catalysis
B: Environmental, 119, 267–272.
33. Yu, C.-L., Jimmy, C. Y., He, H.-B., & Zhou, W.-Q. (2016). Progress in sonochemical fabrica-
tion of nanostructured photocatalysts. Rare Metals, 35(3), 211–222.
34. Franco, F., Pérez-Maqueda, L., & Pérez-Rodrıguez, J. (2004). The effect of ultrasound on
the particle size and structural disorder of a well-ordered kaolinite. Journal of Colloid and
Interface Science, 274(1), 107–117.
35. Ma, Y., Wang, X., Jia, Y., Chen, X., Han, H., & Li, C. (2014). Titanium dioxide-based nano-
materials for photocatalytic fuel generations. Chemical Reviews, 114(19), 9987–10043.
36. Liu, Y., Geng, P., Wang, J., Yang, Z., Lu, H., Hai, J., Lu, Z., Fan, D., & Li, M. (2018). In-situ
ion-exchange synthesis Ag2S modified SnS2 nanosheets toward highly photocurrent response
and photocatalytic activity. Journal of Colloid and Interface Science, 512, 784–791.
37. Ma, L., Xu, L., Xu, X., Zhang, L., & Zhou, X. (2016). Fabrication of SnO2/SnS2 hybrids by
anchoring ultrafine SnO2 nanocrystals on SnS2 nanosheets and their photocatalytic proper-
ties. Ceramics International, 42(4), 5068–5074.
480 Udayabhanu et al.

38. Kumar, G. M., Xiao, F., Ilanchezhiyan, P., Yuldashev, S., & Kang, T. (2016). Enhanced pho-
toelectrical performance of chemically processed SnS 2 nanoplates. RSC Advances, 6(102),
99631–99637.
39. Liu, Y., Mi, X., Wang, J., Li, M., Fan, D., Lu, H., & Chen, X. (2019). Two-dimensional SnS 2
nanosheets exfoliated from an inorganic–organic hybrid with enhanced photocatalytic activ-
ity towards Cr (VI) reduction. Inorganic Chemistry Frontiers, 6(4), 948–954.
40. Wang, S., & Li, J. (2015). Two-dimensional inorganic–organic hybrid semiconductors com-
posed of double-layered ZnS and monoamines with aromatic and heterocyclic aliphatic
rings: Syntheses, structures, and properties. Journal of Solid State Chemistry, 224, 40–44.
41. Giannakas, A., Seristatidou, E., Deligiannakis, Y., & Konstantinou, I. (2013). Photocatalytic
activity of N-doped and N–F co-doped TiO2 and reduction of chromium (VI) in aqueous
solution: An EPR study. Applied Catalysis B: Environmental, 132, 460–468.
42. Fan, S., Li, X., Zhao, Q., Zeng, L., Zhang, M., Yin, Z., Lian, T., Tadé, M. O., & Liu, S. (2018).
Rational design and synthesis of highly oriented copper–zinc ferrite QDs/titania NAE nano-­
heterojunction composites with novel photoelectrochemical and photoelectrocatalytic behav-
iors. Dalton Transactions, 47(36), 12769–12782.
43. Shen, J., Yan, B., Shi, M., Ma, H., Li, N., & Ye, M. (2011). One step hydrothermal synthesis
of TiO2-reduced graphene oxide sheets. Journal of Materials Chemistry, 21(10), 3415–3421.
44. Xiang, Q., Yu, J., & Jaroniec, M. (2011). Preparation and enhanced visible-light photo-
catalytic H2-production activity of graphene/C3N4 composites. The Journal of Physical
Chemistry C, 115(15), 7355–7363.
45. Feng, M., Sun, R., Zhan, H., & Chen, Y. (2010). Lossless synthesis of graphene nanosheets
decorated with tiny cadmium sulfide quantum dots with excellent nonlinear optical proper-
ties. Nanotechnology, 21(7), 075601.
46. Nethravathi, C., Nisha, T., Ravishankar, N., Shivakumara, C., & Rajamathi, M. (2009).
Graphene–nanocrystalline metal sulphide composites produced by a one-pot reaction start-
ing from graphite oxide. Carbon, 47(8), 2054–2059.
47. Wang, K., Liu, Q., Guan, Q.-M., Wu, J., Li, H.-N., & Yan, J.-J. (2011). Enhanced direct elec-
trochemistry of glucose oxidase and biosensing for glucose via synergy effect of graphene
and CdS nanocrystals. Biosensors and Bioelectronics, 26(5), 2252–2257.
48. Jia, L., Wang, D.-H., Huang, Y.-X., Xu, A.-W., & Yu, H.-Q. (2011). Highly durable N-doped
graphene/CdS nanocomposites with enhanced photocatalytic hydrogen evolution from water
under visible light irradiation. The Journal of Physical Chemistry C, 115(23), 11466–11473.
49. Li, Q., Guo, B., Yu, J., Ran, J., Zhang, B., Yan, H., & Gong, J. R. (2011). Highly efficient
visible-light-driven photocatalytic hydrogen production of CdS-cluster-decorated graphene
nanosheets. Journal of the American Chemical Society, 133(28), 10878–10884.
50. Li, Z., Yao, Y., Lin, Z., Moon, K.-S., Lin, W., & Wong, C. (2010). Ultrafast, dry microwave
synthesis of graphene sheets. Journal of Materials Chemistry, 20(23), 4781–4783.
51. Zheng, X., Weng, J., & Hu, B. (2010). Microwave-assisted synthesis of mesoporous CdS
quantum dots modified by oleic acid. Materials Science in Semiconductor Processing, 13(3),
217–220.
52. Liu, X., Pan, L., Lv, T., Zhu, G., Sun, Z., & Sun, C. (2011). Microwave-assisted synthesis of
CdS–reduced graphene oxide composites for photocatalytic reduction of Cr (vi). Chemical
Communications, 47(43), 11984–11986.
53. Tang, L., Wang, Y., Li, Y., Feng, H., Lu, J., & Li, J. (2009). Preparation, structure, and elec-
trochemical properties of reduced graphene sheet films. Advanced Functional Materials,
19(17), 2782–2789.
54. Zhang, H., Lv, X., Li, Y., Wang, Y., & Li, J. (2009). P25-graphene composite as a high perfor-
mance photocatalyst. ACS Nano, 4(1), 380–386.
55. Yuan, X., Wang, H., Wang, J., Zeng, G., Chen, X., Wu, Z., Jiang, L., Xiong, T., Zhang, J., &
Wang, H. (2018). Near-infrared-driven Cr (VI) reduction in aqueous solution based on a MoS
2/Sb 2 S 3 photocatalyst. Catalysis Science & Technology, 8(6), 1545–1554.
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 481

56. Zhou, W., Yin, Z., Du, Y., Huang, X., Zeng, Z., Fan, Z., Liu, H., Wang, J., & Zhang, H. (2013).
Synthesis of few-layer MoS2 nanosheet-coated TiO2 nanobelt heterostructures for enhanced
photocatalytic activities. Small, 9(1), 140–147.
57. Hu, X., Zhu, Q., Wang, X., Kawazoe, N., & Yang, Y. (2015). Nonmetal–metal–semiconductor-­
promoted P/Ag/Ag 2 O/Ag 3 PO 4/TiO 2 photocatalyst with superior photocatalytic activity
and stability. Journal of Materials Chemistry A, 3(34), 17858–17865.
58. Hu, S., Li, Y., Li, F., Fan, Z., Ma, H., Li, W., & Kang, X. (2016). Construction of g-C3N4/Zn0.
11Sn0. 12Cd0. 88S1. 12 hybrid heterojunction catalyst with outstanding nitrogen photofixa-
tion performance induced by sulfur vacancies. ACS Sustainable Chemistry & Engineering,
4(4), 2269–2278.
59. Bokare, A. D., & Choi, W. (2011). Advanced oxidation process based on the Cr (III)/Cr (VI)
redox cycle. Environmental Science & Technology, 45(21), 9332–9338.
60. Gao, W., Liu, W., Leng, Y., Wang, X., Wang, X., Hu, B., Yu, D., Sang, Y., & Liu, H. (2015).
In2S3 nanomaterial as a broadband spectrum photocatalyst to display significant activity.
Applied Catalysis B: Environmental, 176, 83–90.
61. Lv, T., Pan, L., Liu, X., Lu, T., Zhu, G., Sun, Z., & Sun, C.  Q. (2012). One-step synthe-
sis of CdS–TiO 2–chemically reduced graphene oxide composites via microwave-assisted
reaction for visible-light photocatalytic degradation of methyl orange. Catalysis Science &
Technology, 2(4), 754–758.
62. Ebraheem, S., & El-Saied, A. (2013). Band gap determination from diffuse reflectance mea-
surements of irradiated lead borate glass system doped with TiO2 by using diffuse reflectance
technique. Materials Sciences and Applications, 4(05), 324.
63. Tong, T., Wilke, C.  M., Wu, J., Binh, C.  T. T., Kelly, J.  J., Gaillard, J.-F.  O., & Gray,
K. A. (2015). Combined toxicity of nano-ZnO and nano-TiO2: From single-to multinanoma-
terial systems. Environmental Science & Technology, 49(13), 8113–8123.
64. Liu, B., & Aydil, E. S. (2009). Growth of oriented single-crystalline rutile TiO2 nanorods
on transparent conducting substrates for dye-sensitized solar cells. Journal of the American
Chemical Society, 131(11), 3985–3990.
65. Vignesh, K., Priyanka, R., Hariharan, R., Rajarajan, M., & Suganthi, A. (2014). Fabrication
of CdS and CuWO4 modified TiO2 nanoparticles and its photocatalytic activity under visible
light irradiation. Journal of Industrial and Engineering Chemistry, 20(2), 435–443.
66. Yu, S., Kim, Y. H., Lee, S. Y., Song, H. D., & Yi, J. (2014). Hot-electron-transfer enhance-
ment for the efficient energy conversion of visible light. Angewandte Chemie International
Edition, 53(42), 11203–11207.
67. Kim, W., Park, J.  Y., & Kim, Y. (2019). Fabrication of branched-TiO2 microrods on the
FTO glass for photocatalytic reduction of Cr (VI) under visible-light irradiation. Journal of
Industrial and Engineering Chemistry, 73, 248–253.
68. Park, S., Selvaraj, R., Meetani, M. A., & Kim, Y. (2017). Enhancement of visible-light-driven
photocatalytic reduction of aqueous Cr (VI) with flower-like In3+-doped SnS2. Journal of
Industrial and Engineering Chemistry, 45, 206–214.
69. Wang, K., Wu, X., Zhang, G., Li, J., & Li, Y. (2018). Ba5Ta4O15 nanosheet/AgVO3 nanorib-
bon heterojunctions with enhanced photocatalytic oxidation performance: Hole dominated
charge transfer path and plasmonic effect insight. ACS Sustainable Chemistry & Engineering,
6(5), 6682–6692.
70. Yu, J., Zhuang, S., Xu, X., Zhu, W., Feng, B., & Hu, J. (2015). Photogenerated electron res-
ervoir in hetero-p–n CuO–ZnO nanocomposite device for visible-light-driven photocatalytic
reduction of aqueous Cr (VI). Journal of Materials Chemistry A, 3(3), 1199–1207.
71. Deng, S., & Bai, R. (2004). Removal of trivalent and hexavalent chromium with aminated
polyacrylonitrile fibers: Performance and mechanisms. Water Research, 38(9), 2424–2432.
72. Bellú, S., Sala, L., González, J., García, S., Frascaroli, M., Blanes, P., García, J., Peregrin,
J.  S., Atria, A., & Ferrón, J. (2010). Thermodynamic and dynamic of chromium biosorp-
tion by Pectic and lignocellulocic biowastes. Journal of Water Resource and Protection,
2(10), 888.
482 Udayabhanu et al.

73. Tanaka, A., Nakanishi, K., Hamada, R., Hashimoto, K., & Kominami, H. (2013).
Simultaneous and stoichiometric water oxidation and Cr (VI) reduction in aqueous sus-
pensions of functionalized plasmonic photocatalyst Au/TiO2–Pt under irradiation of green
light. ACS Catalysis, 3(8), 1886–1891.
74. Mor, G. K., Varghese, O. K., Wilke, R. H., Sharma, S., Shankar, K., Latempa, T. J., Choi,
K.-S., & Grimes, C. A. (2008). p-Type Cu− Ti− O nanotube arrays and their use in self-­
biased heterojunction photoelectrochemical diodes for hydrogen generation. Nano Letters,
8(7), 1906–1911.
75. Mor, G. K., Prakasam, H. E., Varghese, O. K., Shankar, K., & Grimes, C. A. (2007). Vertically
oriented Ti− Fe− O nanotube array films: Toward a useful material architecture for solar
spectrum water photoelectrolysis. Nano Letters, 7(8), 2356–2364.
76. Jiang, Z., Yang, F., Luo, N., Chu, B. T., Sun, D., Shi, H., Xiao, T., & Edwards, P. P. (2008).
Solvothermal synthesis of N-doped TiO 2 nanotubes for visible-light-responsive photocataly-
sis. Chemical Communications, (47), 6372–6374.
77. Kuang, S., Yang, L., Luo, S., & Cai, Q. (2009). Fabrication, characterization and photoelec-
trochemical properties of Fe2O3 modified TiO2 nanotube arrays. Applied Surface Science,
255(16), 7385–7388.
78. Kongkanand, A., Tvrdy, K., Takechi, K., Kuno, M., & Kamat, P. V. (2008). Quantum dot solar
cells. Tuning photoresponse through size and shape control of CdSe− TiO2 architecture.
Journal of the American Chemical Society, 130(12), 4007–4015.
79. Bessekhouad, Y., Robert, D., & Weber, J.-V. (2005). Photocatalytic activity of Cu2O/TiO2,
Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions. Catalysis Today, 101(3–4), 315–321.
80. Bechinger, C., Oefinger, G., Herminghaus, S., & Leiderer, P. (1993). On the fundamental
role of oxygen for the photochromic effect of WO3. Journal of Applied Physics, 74(7),
4527–4533.
81. Tatsuma, T., Saitoh, S., Ohko, Y., & Fujishima, A. (2001). TiO2− WO3 photoelectrochemi-
cal anticorrosion system with an energy storage ability. Chemistry of Materials, 13(9),
2838–2842.
82. Su, W.-B., Wang, J.-F., Chen, H.-C., Wang, W.-X., Zang, G.-Z., & Li, C.-P. (2003). Novel
TiO2· WO3 varistor system. Materials Science and Engineering: B, 99(1–3), 461–464.
83. Akurati, K.  K., Vital, A., Dellemann, J.-P., Michalow, K., Graule, T., Ferri, D., & Baiker,
A. (2008). Flame-made WO3/TiO2 nanoparticles: Relation between surface acidity, structure
and photocatalytic activity. Applied Catalysis B: Environmental, 79(1), 53–62.
84. Keller, V., Bernhardt, P., & Garin, F. (2003). Photocatalytic oxidation of butyl acetate in vapor
phase on TiO2, Pt/TiO2 and WO3/TiO2 catalysts. Journal of Catalysis, 215(1), 129–138.
85. Georgieva, J., Armyanov, S., Valova, E., Poulios, I., & Sotiropoulos, S. (2007). Enhanced
photocatalytic activity of electrosynthesised tungsten trioxide–titanium dioxide bi-layer coat-
ings under ultraviolet and visible light illumination. Electrochemistry Communications, 9(3),
365–370.
86. Meichtry, J. M., Brusa, M., Mailhot, G., Grela, M. A., & Litter, M. I. (2007). Heterogeneous
photocatalysis of Cr (VI) in the presence of citric acid over TiO2 particles: Relevance of Cr
(V)–citrate complexes. Applied Catalysis B: Environmental, 71(1–2), 101–107.
87. Kong, C., Li, M., Li, J., Ma, X., Feng, C., & Liu, X. (2019). One-step synthesis of Fe 2
O 3 nano-rod modified reduced graphene oxide composites for effective Cr (vi) removal:
Removal capability and mechanism. RSC Advances, 9(36), 20582–20592.
88. Stock, N. L., Peller, J., Vinodgopal, K., & Kamat, P. V. (2000). Combinative sonolysis and
photocatalysis for textile dye degradation. Environmental Science & Technology, 34(9),
1747–1750.
89. Zheng, Y., Zheng, L., Zhan, Y., Lin, X., Zheng, Q., & Wei, K. (2007). Ag/ZnO heterostructure
nanocrystals: Synthesis, characterization, and photocatalysis. Inorganic Chemistry, 46(17),
6980–6986.
90. Lu, W., Wei, Z., Gu, Z., Liu, T., Park, J., Park, J., Tian, J., Zhang, M., Zhang, Q., Gentle, T.,
III, Bosch, M., & Zhou, H.-C. (2014). Chemical Society Reviews, 43, 5561.
18  Recent Advances in Nanostructured Materials for Detoxification of Cr(VI… 483

91. Fu, H.-R., Kang, Y., & Zhang, J. (2014). Highly selective sorption of small hydrocar-
bons and photocatalytic properties of three metal–organic frameworks based on Tris (4-(1
H-imidazol-1-yl) phenyl) amine ligand. Inorganic Chemistry, 53(8), 4209–4214.
92. Zhao, J., Dong, W.-W., Wu, Y.-P., Wang, Y.-N., Wang, C., Li, D.-S., & Zhang, Q.-C. (2015).
Two (3, 6)-connected porous metal–organic frameworks based on linear trinuclear [Co 3
(COO) 6] and paddlewheel dinuclear [Cu 2 (COO) 4] SBUs: Gas adsorption, photocatalytic
behaviour, and magnetic properties. Journal of Materials Chemistry A, 3(13), 6962–6969.
93. Zhang, C., Ai, L., & Jiang, J. (2015). Solvothermal synthesis of MIL–53 (Fe) hybrid mag-
netic composites for photoelectrochemical water oxidation and organic pollutant photodeg-
radation under visible light. Journal of Materials Chemistry A, 3(6), 3074–3081.
94. Zhang, C.-F., Qiu, L.-G., Ke, F., Zhu, Y.-J., Yuan, Y.-P., Xu, G.-S., & Jiang, X. (2013). A
novel magnetic recyclable photocatalyst based on a core–shell metal–organic framework Fe
3 O 4@ MIL-100 (Fe) for the decolorization of methylene blue dye. Journal of Materials
Chemistry A, 1(45), 14329–14334.
95. Wang, H., Yuan, X., Wu, Y., Zeng, G., Chen, X., Leng, L., Wu, Z., Jiang, L., & Li, H. (2015).
Facile synthesis of amino-functionalized titanium metal-organic frameworks and their
superior visible-light photocatalytic activity for Cr (VI) reduction. Journal of Hazardous
Materials, 286, 187–194.
96. Liang, R., Shen, L., Jing, F., Wu, W., Qin, N., Lin, R., & Wu, L. (2015). NH2-mediated
indium metal–organic framework as a novel visible-light-driven photocatalyst for reduction
of the aqueous Cr (VI). Applied Catalysis B: Environmental, 162, 245–251.
97. Zhao, H., Xia, Q., Xing, H., Chen, D., & Wang, H. (2017). Construction of pillared-layer
MOF as efficient visible-light photocatalysts for aqueous Cr (VI) reduction and dye degrada-
tion. ACS Sustainable Chemistry & Engineering, 5(5), 4449–4456.
98. Chen, W., Yang, Z., Xie, Z., Li, Y., Yu, X., Lu, F., & Chen, L. (2019). Chen, Benzothiadiazole
functionalized D–A type covalent organic frameworks for effective photocatalytic reduction
of aqueous chromium (VI). Journal of Materials Chemistry A, 7(3), 998–1004.
99. Cheng, Y., Yan, F., Huang, F., Chu, W., Pan, D., Chen, Z., Zheng, J., Yu, M., Lin, Z., &
Wu, Z. (2010). Bioremediation of Cr (VI) and immobilization as Cr (III) by Ochrobactrum
anthropi. Environmental Science & Technology, 44(16), 6357–6363.
100. Kim, D.-H., Park, S., Kim, M.-G., & Hur, H.-G. (2014). Accumulation of amorphous Cr
(III)–Te (IV) nanoparticles on the Surface of Shewanella oneidensis MR-1 through reduction
of Cr (VI). Environmental Science & Technology, 48(24), 14599–14606.
101. Belchik, S. M., Kennedy, D. W., Dohnalkova, A. C., Wang, Y., Sevinc, P. C., Wu, H., Lin,
Y., Lu, H.  P., Fredrickson, J.  K., & Shi, L. (2011). Extracellular reduction of hexavalent
chromium by cytochromes MtrC and OmcA of Shewanella oneidensis MR-1. Applied and
Environmental Microbiology, 77(12), 4035–4041.
102. Middleton, S. S., Latmani, R. B., Mackey, M. R., Ellisman, M. H., Tebo, B. M., & Criddle,
C.  S. (2003). Cometabolism of Cr (VI) by Shewanella oneidensis MR-1 produces cell-­
associated reduced chromium and inhibits growth. Biotechnology and Bioengineering, 83(6),
627–637.
103. Daulton, T.  L., Little, B.  J., Lowe, K., & Jones-Meehan, J. (2002). Electron energy loss
spectroscopy techniques for the study of microbial chromium (VI) reduction. Journal of
Microbiological Methods, 50(1), 39–54.
104. Wang, Y., Sevinc, P. C., Belchik, S. M., Fredrickson, J., Shi, L., & Lu, H. P. (2013). Single-­
cell imaging and spectroscopic analyses of Cr (VI) reduction on the surface of bacterial cells.
Langmuir, 29(3), 950–956.
105. Dong, G., Wang, Y., Gong, L., Wang, M., Wang, H., He, N., Zheng, Y., & Li, Q. (2013).
Formation of soluble Cr (III) end-products and nanoparticles during Cr (VI) reduction by
Bacillus cereus strain XMCr-6. Biochemical Engineering Journal, 70, 166–172.
106. Kim, D.-H., Kanaly, R.  A., & Hur, H.-G. (2012). Biological accumulation of tellurium
nanorod structures via reduction of tellurite by Shewanella oneidensis MR-1. Bioresource
Technology, 125, 127–131.
484 Udayabhanu et al.

107. Zhang, H.-K., Lu, H., Wang, J., Zhou, J.-T., & Sui, M. (2014). Cr (VI) reduction and Cr (III)
immobilization by Acinetobacter sp. HK-1 with the assistance of a novel quinone/graphene
oxide composite. Environmental Science & Technology, 48(21), 12876–12885.
108. Chen, Y., & Gu, G. (2005). Preliminary studies on continuous chromium (VI) biologi-
cal removal from wastewater by anaerobic–aerobic activated sludge process. Bioresource
Technology, 96(15), 1713–1721.
109. Das, S., Mishra, J., Das, S. K., Pandey, S., Rao, D. S., Chakraborty, A., Sudarshan, M., Das,
N., & Thatoi, H. (2014). Investigation on mechanism of Cr (VI) reduction and removal by
Bacillus amyloliquefaciens, a novel chromate tolerant bacterium isolated from chromite mine
soil. Chemosphere, 96, 112–121.
110. Liu, G., Yang, H., Wang, J., Jin, R., Zhou, J., & Lv, H. (2010). Enhanced chromate reduction
by resting Escherichia coli cells in the presence of quinone redox mediators. Bioresource
Technology, 101(21), 8127–8131.
111. Li, B., Pan, D., Zheng, J., Cheng, Y., Ma, X., Huang, F., & Lin, Z. (2008). Microscopic
investigations of the Cr (VI) uptake mechanism of living Ochrobactrum anthropi. Langmuir,
24(17), 9630–9635.
112. Pan, X., Liu, Z., Chen, Z., Cheng, Y., Pan, D., Shao, J., Lin, Z., & Guan, X. (2014).
Investigation of Cr (VI) reduction and Cr (III) immobilization mechanism by planktonic cells
and biofilms of Bacillus subtilis ATCC-6633. Water Research, 55, 21–29.
Chapter 19
Metal Nitrides and Graphitic Carbon
Nitrides as Novel Photocatalysts
for Hydrogen Production
and Environmental Remediation

Sudesh Kumar, Kakarla Raghava Reddy, Ch. Venkata Reddy,


Nagaraj P. Shetti, Veera Sadhu, M. V. Shankar, Vasu Govardhana Reddy,
A. V. Raghu, and Tejraj M. Aminabhavi

19.1  Introduction

In the twenty-first century, humans are facing serious problems to provide renew-
able and clean energy to our modern society. Photocatalysis has been studied and is
probable to gross an abundant influence on eco-friendly emissions and renewable
energy. Photocatalytic hydrogen generation technology from water is the greatest
encouraging method to grasp an economy of hydrogen due to usage of solar energy
(its clean and enduring energy source); it is an ecologically harmless method

S. Kumar
Department of Chemistry, Banasthali Vidyapeeth, Vanasthali, Rajasthan, India
K. R. Reddy (*)
School of Chemical and Biomolecular Engineering, The University of Sydney,
Sydney, NSW, Australia
C. V. Reddy (*)
School of Mechanical Engineering, Yeungnam University, Gyeongsan, South Korea
N. P. Shetti
Department of Chemistry, K. L. E. Institute of Technology, Hubli, Karnataka, India
Visvesvaraya Technological University, Belgaum, Karnataka, India
V. Sadhu
School of Physical Sciences, Kakatiya Institute of Technology and Science (KITS),
Warangal, Telangana, India
M. V. Shankar
Nanocatalysis and Solar Fuels Research Lab, Department of Materials Science and
Nanotechnology, Yogi Vemana University, Kadapa, Andhra Pradesh, India

© Springer Nature Switzerland AG 2021 485


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_19
486 S. Kumar et al.

without unwanted products and contaminants, and solar energy can be converted
into a storable energy form [1].
Metal nitrides are a significant class of developing materials in optimum cases
and may syndicate the benefits of oxides and other materials owing to their enhanced
electrical conductivity and good corrosion resistance [2]. Generally, in nitride,
nitrogen is combined with a low electronegativity element such as B, Si, and metals.
Nitride is a nitrogen-containing compound having an oxidation state of −3 with a
wide range of properties along with immense applications, such as optoelectronic
devices [3], solid-state gas sensors [4], batteries, and microelectronic devices [5],
Fuel cells [6], corrosion resistance coatings [7], hydrogen storage [8], and semicon-
ductive properties [9, 10], respectively.
The amount of sun energy conveyed yearly is around 3 × 1024 J of vitality to the
earth’s surface [11]. TiO2 has been the widely utilized photocatalyst since its early
development in 1972 by Fujishima and Honda [12]. Under the bright irradiation,
TiO2 impetus is energized, which accounts less than 5% of the entire sun-powered
range. This has created greater interests among scientists to create novel materials
with shorter bandgap vitality (Eg) in order to improve the reaction time to the light
photons. Different altered TiO2 or non-TiO2 catalysts have been employed as the
light-determined photocatalysts [13, 14]. Recently, metal nitrides and graphitic car-
bon nitride (g-C3N4) have been turned out to be a wonderful material in photocataly-
sis and for hydrogen production [15].
Since then metal nitrides and transition metal nitrides are used as photocatalyst
detectors because of their wide applications in energy storage [16], hydrogen pro-
duction [17], sustainable organic semiconductor, and environmental purification
[18, 19]. There is thus a need to discuss the preparation of metal nitride, its different
types, synthesis, mechanism, surface phenomenon, and their applications in energy
and environmental divisions. The present chapter is an attempt in this direction that
will address all aspects of metal nitrides as novel alternatives to conventional
photocatalysis.

V. G. Reddy
Department of Chemistry, Yogi Vemana University, Kadapa, Andhra Pradesh, India
A. V. Raghu
Department of Chemistry, Faculty of Engineering and Technology, Jain (Deemed-to-be
University), Bangalore, Karnataka, India
T. M. Aminabhavi
Sonia College of Pharmacy, Dharwad, Karnataka, India
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 487

19.2  Classification of Nitrides

Nitrides are synthesized by two principal methods of which the first one being the
direct reaction approach at elevated temperature (e.g., preparation of calcium
nitride).

Ca 3 N 2 .3Ca + N 2 → Ca 3 N 2 A

The second approach is the loss of NH3 gas evolution via heat degradation of a
metal amide (e.g., synthesis of barium amide). Nitrides are formed through the sur-
face reinforcing process when ammonia is heated at high temperatures of 500–550 °C
for 5–100 h. An alternative method, which deals with the chemically reducing metal
oxide or halide in the presence of inert gas (nitrogen) to form nitrides, e.g., synthesis
of aluminum nitride.

19.2.1  Ionic Nitrides

Lithium (Li) is an alkali metal ready to form a nitride, albeit complete alkali earth
metals to form nitrides, e.g., M3N2. The mixes comprising the cations of N3− anions
and metals experience hydrolysis (response with water) and give smelling salts and
metal hydroxide. The constancy of ionic nitrides displays varied temperature ranges,
e.g., Mg3N2 [20] that degrades beyond 270 °C (520 °F), although Be3N2 [21] melts
at 2200 °C, but without breakdown.

19.2.2  Interstitial Nitrides

The interstitial nitrides can be comparable in direction of the interstitial carbides


where N2 atoms possess the interstitial sites. The typical equations of such type of
nitrides are M4N [22], M2N [23], and MN despite the fact that their stoichiometry
variation. These mixes have great melting focuses, are greatly rigid, and are nor-
mally murky materials that have metallic structure and extraordinary conductivities.
In general, they are ordinarily arranged by pyrolyzing the metal in smelling salts at
about 1200 °C. These are synthetically inactive and known couples of responses.
The most trademark response is hydrolysis, which is typically moderate, to deliver
smelling salts or nitrogen gas with the liberation of H2 gas.

2 VN + 3H 2 SO 4 → V2 ( SO 4 )3 + N 2 + 3H 2

Under their manufactured dormancy and the resistance capability for higher tem-
peratures, interstitial nitrides are very beneficial for including their usage as caul-
drons and vessels for elevated thermal reactions.
488 S. Kumar et al.

19.2.3  Covalent Nitrides

The covalent parallel nitrides have an extensive variety of properties relying upon
the component to which nitrogen is fortified. A few cases of covalently bonded
nitrides include boron nitride (BN), disulfurdinitride (S2N2), phosphorus nitride
(P3N5), cyanogen ((CN)2), and tetrasulfurtetranitride (S4N4). The details of covalent
nitrides such as carbon, boron, and sulfur are mentioned [24, 25].

19.2.4  Boron Nitrides

Since boron and nitrogen remain collective containing a similar in outer most shell
electrons as two fortified carbon atoms, the boron nitride (BN) is found to possess
the same electronic structure with the essential carbon [26]. However, BN has two
basic structures that are comparable to two types of carbon graphite and jewel.
Graphite has a hexagonal shape and a layered planar structure with six-membered
rings of exchanging B and N atoms are arranged in a way that a B particle in one
layer is found straightforwardly finished atoms in the neighboring layer. Conversely,
the graphite’s progressive hexagonal layers are balanced with the goal that every
carbon molecule is specifically over an interstice (opening) in a contiguous layer
and straightforwardly finished a carbon particle of exchange layered. Hexagonal-­
shaped boron nitride is fabricated via warming boron trichloride precursor (BCl3)
under over abundance of smelling salts at 750 °C.
The physicochemical characteristics of BN are not quite the same as that of
graphite. Even though both are tricky entities, BN is vapid and is a decent cover
(though graphite is dark and electrically conductive) and is largely steady artificially
compared to graphite. BN responds just like basic fluorine, F2 (framing the items
BF3 and N2), and HF (delivering NH4BF4). Warming BN to 1800 °C under large
weight (85,000 air; the weight adrift-levelled environment) within the sight of a
metal salt or soluble base earth metal impetus results in a precious stone (cubic) like
BN. Just similar to a precious stone type of carbon, cubic boron nitride is to a great
degree hard [27].

19.2.5  Sulfur Nitrides

Many varieties of covalent binary nitrides can be obtained from sulfur; however, the
utmost exciting nitrides are tetrasulfurtetranitride (S4N4) [28] and disulfurdinitride
(S2N2) due to these two nitrides are the precursors to polythiazyl ((SN)x), unfamiliar
polymer. Polythiazyl is unusual because of its properties which are associated only
with metals, and it is composed of solely two nonmetals. The S4N4 preparation
involves bubbling of ammonia gas into a heated S2Cl2 solution (50  °C) either in
CCl4 or C6H6.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 489

Tetrasulfurtetranitride generally creates thermochromic crystals; their color is


sensitive to heat. It appears colorless at the temperature of −190 °C, orange at 25 °C,
and above 100 °C. S4N4 crystal particles have good stability to oxygen, but they will
explode in reaction to friction or shock. Crystals have a cage-shaped arrangement
containing four planer N atoms and two S atoms present on either side of the plane.
When the S4N4 vapor is filled through the silver wool about 250–300 °C and low
pressure (<1.0 mmHg), an unstable dimer, S2N2, is formed. The dimer possesses a
square shape that has repetition of atoms of nitrogen and sulfur. S2N2 is sensitive to
friction or shock. Polythiazyl [(SN)x] forms through the ring-opening mechanism.
This appears in bronze color and has a metallic luster, metal-like thermal, as well as
electrical conductivity and has superconductivity around 0.26 K [29].

19.3  Graphitic Carbon Nitride (g-C3N4)

Graphitic carbon nitride (g-C3N4) is the utmost stable allotrope of carbon nitrides at
ambient atmosphere, but it also has rich surface properties that are suitable for
numerous applications, due to the existence of basic surface sites. The ideal g-C3N4
contains solely a gathering of C–N bonds without electron localization in the π-state
(this material is a π-conjugated polymer) [30]. Likewise, due to the occurrence of
hydrogen and to the fact that nitrogen has one more electron than carbon, g-C3N4
has rich surface properties. Furthermore, its extraordinary thermal stability (it is
stable up to 600 °C in air) and hydrothermal stability (it is insoluble either in acidic,
neutral, or basic solvents) enables the material to function either in liquid or gaseous
environments and, at elevated temperatures, potentiating its wide applications [31].

19.3.1  Fabrication of Bulk β-C3N4

β-C3N4 can be incorporated by mechanochemical process. In an argon climate,


exceedingly unadulterated graphite powders down to an undefined nanoscale esti-
mate through a ball process taken after argon cleansing. NH3 gas environment is
required for the addition of graphite powder. In the wake of processing, a nano-sized
piece-like structure of β-C3N4 has been found [32]. Ball or powder crashes cause
breakage and joining of the graphite fine powder and the reactants again and again.
Plastic wrap of powdered graphite happens as a result of the shear bunches which
break down to sub-grains which are isolated by low-point granular limits, further-
more preparing the sub-grain to evaluate until nano-sized sub-grains shape. The
high weight and extraordinary development propel reactant partition of NH3 parti-
cles into monoatomic nitrogen on the carbon split surface. The nano-sized carbon
powder acts liberally not the same as its mass material as a consequence of the
estimation of particle and surface zone, producing a nanoscale form of carbon to
adequately interact with free N atoms, gives β-C3N4 fine powder [33].
490 S. Kumar et al.

β-C3N4 nanorods can also be synthesized through alternative methods. The


powder-­like compound is thermally toughened with NH3 gas stream, and single-­
crystal β-C3N4 nanorods are prepared. The span of the nanorods is dictated by the
time and temperature of warm strengthening. These nanorods become fastened in
their pivot course that the breadth heading and having finishes line hemispherical
nanorods cross-area demonstrates that their prismatic segment morphology. And it
was found that they contain shapeless stages, at 450 °C temperature for 3 h in NH3
environment, and the measure of the indistinct stage decreased to nothing. When
compared to nanotubes, these nanorods exist as twinned and thick. Integrating
nanorods by warm tempering gives a powerful, minimal effort and high return strat-
egy for the union of nanorod-shaped single crystals. Instead of framing a nanopow-
der, the β-C3N4 compound can on the other hand be shaped in thin undefined movies
by either stun wave pressure innovation, pyrolysis of antecedents with excess nitro-
gen composition, diode sputtering, solvothermal planning, beat laser removal, or
particle implantation [34].

19.3.2  Challenges in the Processing of β-C3N4

The strategy and amalgamation of encircled carbon nitride have been represented.
The nitrogen union of complex has a tendency to be underneath the idyllic ensemble
for C3N4, in light of the low thermodynamic consistent quality concerning the seg-
ments C and N2, which appeared through favorably estimating enthalpy game plan.
The mechanical utilization of nanopowder is amazingly restricted by the extensive
mix cost nearby troublesome systems for creation that causes a low yield [35–37].

19.3.3  Applications of β-C3N4

β-C3N4 has many applications in various fields such as tribology, electronic engi-
neering, optical engineering, and wear-resistant coating [38]. Industrialization inno-
vations prompt many advantages for people yet additionally cause numerous natural
issues, for example, air and water contamination. To limit or moderate the defile-
ments, metal nitride assumes a crucial part. Metal nitride photograph impetuses are
carbon nitride, carbon nitride edifices, ruthenium carbon nitride complex, graphitic
carbon nitride, graphitic carbon nitride-silver, titanium oxide complex, and sensitiz-
ers [39]. To be sure, lately during the photograph catalysis, metal nitride has pulled
in consideration as the perfect photograph excitation material to devastate a huge
range of natural mixes, including CO2 lessening, band hole structure, development
of sub-atomic oxygen, counterfeit photosynthetic get-together, hydrogen advance-
ment response, organ catalysis for Friedel-Crafts response, water part, oceanic con-
tamination, obvious light enhancer, haloalkanes, aromatics, alkanes, bug sprays,
pesticides, and surfactants.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 491

The photocatalysis process is an extremely powerful strategy because of its sim-


ple and viable execution. The drawback of utilization of metal oxide like photocata-
lytic material is its retention in the UV area and quick recombination of electrons
[40, 41]. Numerous strategies have been utilized to take care of these issues. The
primary technique is doping with the metal, and the second is the pairing of another
semiconductor [42–44]. Polymer-like semiconductor materials (g-C3N4) are mea-
sured as competent photocatalyst under noticeable light which is used as a without
metal photo driving force because of its abundance, nontoxicity, and moderate vital-
ity band hole (2.7 eV), has pulled in thought in the photocatalysis field, for instance,
sunlight based situated imperativeness change, hydrogen age, and condition decon-
tamination [45–47].
Polymer-like semiconductor materials (e.g., g-C3N4) are favored because of their
high steadiness, minimal effort, and controllable surface. Accordingly, polymer-like
semiconductors can be considered as another material for sunlight-based vitality
and natural applications. The downsides for utilizing like business photocatalyst: a
little surface region, a high recombination rate of the electron-gap sets, and an
absence of retention over 460 nm. Scientists have beaten these disadvantages with
three strategies. The controlling surface of polymer-type semiconductors, which
expands adsorption capacity and photocatalytic movement, is one of them [48, 49].
Another is to expand the assimilation scope of polymer-type semiconductors via
altering their band structure by means of elemental doping [50, 51]. The third one is
to expand the retention scope of polymer-type semiconductors by modifying their
electronic structure by coupling with other semiconducting photocatalysts, for
example, Bi2WO6, BiOCl, TiO2, WO3, TaON, and so on.
When the frequency of the photon and the metal’s plasma oscillation frequency
is consistent together, it will produce incident light with resonance and strong
absorption, which is called local SPR phenomenon. The deposition of Ag onto the
g-C3N4 surface through photo-deposition followed by precipitation and/or calcina-
tion process improved the photocatalytic activity [52]. However, the Ag particles
having high particle size distribution on the substrate’s surface are not homoge-
neous. However, they agglomerate in some areas owing to the variability in the
treatment method and/or irradiation. Utilizing a polymer as a photocatalyst may
speak to a critical advance toward fake photosynthesis since the diminishing and
oxidizing focuses inside photosynthetic living beings are biopolymer [53]. It is
notable that the p-conjugated polymer’s electronic properties can be altered by mix-
ing dopant [54] and copolymerization [55].
Lately, Wang and associates blended [56] a carbon-rich C3N4 photocatalyst with
a high noticeable light movement for hydrogen creation by copolymerization. This
is while immobilizing g-C3N4 on the exterior part of graphene delivered a layered
hybrid with having enhanced conductivity, and electro-synergist execution [57]. In
this way, changing the electronic constitution of the carbon nitride polymer (CNP)
is required to grow high viable polymeric photocatalytic materials with high effi-
ciency. Nitrogen-doped semiconductor was reported to possess notable photocata-
lytic activity in visible lights [58].
492 S. Kumar et al.

19.3.4  Adsorption Kinetics of Nitrides

The adsorption kinetics of metal salt/metal/g-C3N4 can be well fitted through a


pseudo-second-order equation, according to which.

t 1 t
= 2 + e
qt kq e q

Here, k relates to rate constant [g (mg/min)], qe corresponds to equilibrium


adsorption capacity (mg/g), and qt assigns to a concentration of dye (mg/g) adsorbed
during time t. From the plot drawn between t/qt and t, qe and k can be estimated from
the linear fit of the plot. One can decide the estimations of qe and k from the slant
and block of the fitted line, separately. Connection coefficient (R2) estimation related
to linear fit is close solidarity (0.9999), showing relevance to pseudo-second-arrange
display regarding adsorption kinetics which is stated above.

19.4  V
 isible Light Catalytic Reduction of Pb Ions
with Nitride Catalysts

The photocatalytic performance of graphene, carbon nitride, and metal oxide-


doped g-C3N4 was evaluated by its capability to photocatalytically reduce Pb ions
in visible light conditions. Studies were carried out by a batch reactor having a
horizontal cylinder-like circular shape. The 150 W blue fluorescent light was used
to illuminate the photocatalyst, and the catalytic reactor was protected with a UV
cutoff filter. In the photocatalytic experiment, the photocatalyst was dispersed into
300 mL of lead nitrate solution (100 mg/L). This reaction was performed isother-
mally at room temperature, and the aliquots were removed periodically at various
time intervals for an hour. The concentration of Pb was determined using atomic
absorption. The adsorption efficiency of Pb ions is calculated with the following
equation:

%adsorption efficiency of Pb = ( Co − C ) / Co × 100

where Co and C are the initial and the remaining concentrations of Pb in aqueous
solution.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 493

19.4.1  F
 ormation of Graphitic Carbon Nitride
Hybrid Photocatalysts

Most regularly g-C3N4 is set up by a straightforward course by receiving nitrogen-­


rich natural particles as a single-source precursors under heat treatment. g-C3N4 is
formed by immediate warming of melamine in a semi-shut framework [59]. The
cyanamide or dicyandiamide polycondense into g-C3N4 under the nitrogen gas
stream. The warm change of guanylurea dicyanamide to g-C3N4 enables it to solid-
ify to the material with high oxygen content [60]. g-C3N4 is a metal-free, novel, and
unique photo-responsive semiconducting material having an optical band hole of
Eg = 2.7 eV. The g-C3N4 has a photocatalytic movement to generate hydrogen/oxy-
gen amid the water part at a noticeable light intensity. g-C3N4 has a decent execution
regarding photo-decomposition of natural effluents, which depicts its capability in
photocatalysis. The use of g-C3N4 in photocatalysis has increased extensive logical
consideration due to its remarkable optical, electronic, and reactant properties.
Meanwhile, it additionally can be blended from accessible forerunners, for exam-
ple, melamine, dicyandiamide, cyanamide, urea thiourea, and so forth [61].
g-C3N4 experiences the burdens of a low particular surface territory and lower
quantum proficiency, which has constrained its photocatalytic efficiency. The pri-
mary strategy is to utilize metal or nonmetal component doping to enhance the
photoactivity of g-C3N4, (e.g., doping of Fe-, N-, P-, B-, Cu-, and/or S elements with
g-C3N4), which depends on vitality band designing. The second path is to shape
composite impetuses with a moment photocatalyst, for example, TiO2/g-C3N4 [62],
graphene/g-C3N4 [63], SmVO4/g-C3N4 [64], TaON/g-C3N4 [65], N-TiO2/g-C3N4
[66], Cu2O/g-C3N4 [67], SnS2/g-C3N4 [68], and so forth. These composite photo-
catalysts are mainly created because of the coupling impact. Copolymerization with
the nitrogen precursors along with protonation is the third experience. Nevertheless,
the three previously mentioned strategies, expanding the surface area of g-C3N4, has
likewise been viable, as a photocatalytic response can happen at first glance close
by the photocatalytically dynamic focus. Therefore, the photocatalyst covering on
the latent help surface still has the capacity of photocatalysis to corrupt pollutants.
Mesoporous g-C3N4 can be obtained from nano-SiO2 as a hard layout having
extensive surface range and an impressively higher photoactivity over mass of
g-C3N4 in the debasement of 4-chlorophenol. The core/shell nanospheres of SiO2/g-­
C3N4 utilizing a warming technique to strengthen the silicon dioxide nanoparticles
blend as well as liquid cyanamide under the nitrogen atmosphere. Photocatalytic
exercises related to SiO2/g-C3N4 composites demonstrated the most elevated move-
ment with 94.3% RhB transformation after 150 min that is 3.5 times that of pure
g-C3N4. The SiO2-changed g-C3N4 photocatalyst demonstrated that g-C3N4 with
SiO2 as a help accomplishes a surface-active territory. Adsorption of poisons gets
advanced and improves the photocatalytic exercises under noticeable light illumina-
tion. Initiated carbon has an all-around created pore structure with an extensive
surface zone and solid adsorption limit and is generally utilized as an adsorbent and
impetus bolster. What’s more, enacted carbon-rich surface –COOH group
494 S. Kumar et al.

gatherings, phenolic hydroxyl gatherings, carbonyl gatherings, lactone gatherings,


amide gatherings, and so forth can artificially respond with polymers to shape half
and half composites. Moreover, enacted carbon is naturally inexhaustible and eco-
nomical. In any case, some reports are utilizing enacted carbon for blending g-C3N4/
AC hybrid photocatalyst.

19.4.2  Synthesis Strategies of g-C3N4-Based Photocatalysts

g-C3N4 is synthesized by the thermal condensation of nitrogen-rich precursors with


a tri-s-triazine ring structure such as cyanamide, dicyandiamide, urea, or thiourea,
resulting in a graphene-like structure after exfoliation (Fig. 19.1) [39].

19.4.3  P
 hotocatalysis Based on Transition Metal
Nitrides (TMNs)

The transition metal nitrides (TMNs) are intriguing dynamic materials for the
anodes inferable from their important physical properties, such as high softening
focuses. TMNs have excellent mechanical and concoction characteristics, resulting
in unlocking of entryway related to business uses of TMCs and TMNs in appara-
tuses for cutting rotors inside gas turbines, for defensive coatings inside combina-
tion reactors, etc. [69]. The fundamental characteristics of a few particular transition
metal carbides/nitrides are given [70].

Fig. 19.1  Schematic illustration of the synthesis process from the possible precursors of g-C3N4.
Reproduced with permission from [39]; copyright (2016), the American Chemical Society
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 495

Linear thermal
Young’s Heat expansion Electrical
Sample Density Melting modulus conductivity coefficient (10−6) resistivity
name (gm−3) point (°C) (GPa) (W m−1 K−1) K−1 (μΩ)
TiN 5.39 3050 420 29 9.9 27
ZrN 7.32 3000 460 11 7.8 24
HfN 13.83 3330 380 11 8.5 27
VN 6.04 2350 380 11 10.8 65
NbN 8.16 – 360 3.8 10.2 60
TaN 15.9 – – – 8.0 –
CrN 6.14 – 450 11.7 – 640

Transition metal group IVB-VIB nitrides and carbides are referred to “interstitial
alloys.” These are generally prepared by incorporating atomic C or N atoms at inter-
stitial positions of original metals Fe, Co, and Ni which form iron carbide or nitride
[71], CoN [72], and NiN, respectively. It is designated that primary transition metals
have dominant MX- and M2X-type chemical structures, while in later transition,
metal structure is modified to M3X [73].
In the following section, oxygen evolution reaction (OER), hydrogen evolution
reaction (HER), methanol oxidation reaction (MOR), dye-sensitized solar cell
(DSSC), chemical vapor deposition (CVD), chemical bath deposition (CBD),
metal-organic framework (MOF), and atomic layer deposition (ALD) have been
discussed. Vanadium nitride (VN) is synthesized by magnetron sputtering. It finds
application in lithium batteries [74]. VN synthesized by ammonolysis finds use in
supercapacitors [75]. VN is found to exist as nanoparticles when synthesized from
the above-described manner, whereas VN’s synthesis through ammonia reduction
gives nanosheets, which are also used in supercapacitors [76]. TiN is synthesized
via CVD and appears as a film [77]. NixCo2x(OH)6x/TiN and MnO2/TiN hybrid are
synthesized through anodization/ammonia/electrodeposition with a morphology of
nanotube arrays used as supercapacitor [78], TiN/CNT hybrid synthesis occurs via
hydrothermal/ammonia reduction, providing nanotubes which are further used as
DSSCs [79]. TiN@C composite synthesis through hydrothermal/ammonia reduc-
tion produces nanowires [80]. TiN@GNS hybrids find use in DSSC [81]. WN is
synthesized by CVD in the form of films [82]. Mo2N@RGO hybrids were synthe-
sized in the form of nanoparticles by bio-template. Mo2N@GNS composite is syn-
thesized in the form of nanosheets and used in Li-ion batteries [83].

19.5  Metal Nitride Fluorides

The metal nitride fluoride compounds have been a new area for the chemistry [45].
Nitride fluorides are considered to be pseudo-oxides as one N3− and one F− ion
replace two O2− anions in an oxide analog, for example, Mg2NF is an analog for
MgO. Metal nitride fluoride is said to exhibit crystal chemistry which is analogous
to their relative oxides.
496 S. Kumar et al.

19.5.1  Single-Metal Nitride Fluoride

The nitride fluoride compounds were first observed by Lavealle [84]. He reported
some decomposition products such as TcNF and ReNF.  TcNF had a hexagonal
structure with a = 5.98 Å and c = 5.99 Å, and ReNF had a tetragonal structure with
a = 5.88 Å and c = 13.00 Å. The synthesized phase reaction of ZrN powder and
NH4F reflux at a higher temperature. The bandgaps when compared to ZnO are
found to be small (3.2 eV). The studies for photocatalysis were revealed. Both zinc
oxynitride and pure Zn2NF were tested to generate hydrogen using uncontaminated
water. Zn2NF is reported to display more photoactivity than zinc oxynitride fluoride
system. From all the comparisons made, it was concluded how nitrogen changes the
bandgap, thus changing physical and chemical properties.
The synthesis of Zn2NF along with some properties has been reported [85]. The
synthesis started with the reaction of Zn3N2 which was used as a precursor along
with ZnF2. Their ratio was 1:1. The first step was the synthesis of phase through
thermal treatment. The structure was determined through the powder XRD process.
Based on electronic spectra, the characteristics were also explored. A decrease of
bandgap in the phase was observed as a result of nitrogen content in the material.
The bandgap values are 2.8 eV (tetragonal) and 2.7 eV (orthogonal), which was a
carry forward of an earlier work by Anderson in 1970. Recent studies revealed the
synthesis and analysis of Mg3NF3 and Mg2NF. Mg3N2 and MgF2 were allowed to
react at temperature 1050–1150 °C, and the structures were analyzed quantitatively
by XRD and neutron diffraction techniques. The Mg3NF3 is a cube and has a similar
rock salt structure as that of MgO and a vacant cation site. The Mg here is octahe-
dral and is connected to two nitride anions and four fluoride anions. Mg2NF is
tetragonal and has a structure relatable to LiFeO2. Metal nitride fluorides are grouped
broadly among two categories [86], which are single-metal nitride fluorides and
bimetallic nitride fluorides.
The metallic compounds in most of the metal nitride fluorides are Mg, Ca, Sr,
and Ba [87]. Anderson in 1970 reported the synthesis of magnesium nitride fluoride
in the powder phase [88]. The structure of the powder phases was determined with
the help of XRD, and later on the structure was compared to the rock salt structure
of MgO [89]. Mg3NF3 showed similarity in the structure of MgO, but in the case of
Mg3NF3, a cation vacancy was observed [83], where magnesium is seen coordinated
to the four fluoride ions and two nitride ions octahedrally. Nitride and fluoride are
placed at an equal distance of 2.108 Å.
Nitrogen is in octahedral coordination to six cations, and fluorine is centered in
a square configuration attached to four cations [90]. The production of three sys-
tems of metal nitride fluoride: Ca2NF, Sr2NF, and Ba2NF. Sr2NF and Ba2NF were
made by using metal and metal fluoride (molar ratio of 3:1) under nitrogen gas and
a 4:1 molar ratio for Ca2NF. The mixtures were treated with an argon flow at 1000 °C
for 24 h, followed by a nitrogen flow for 24 h at 1000 °C, 950 °C, and 700 °C for
calcium, strontium, and barium, respectively. The Ca2NF product had a light yellow
color, and the colors of the other two phases were not described. The obtained prod-
ucts were analyzed using Debye-Scherrer powder X-ray diffraction, which
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 497

Table 19.1  The structures of metal nitride fluoride of alkali earth metals
Oxide analog (rock salt) Tetragonal Doubled cubic Rock salt Layered hexagonal
MgO L-Mg2NF H-Mg2Nf
CaO L-Ca2NF Ca2NF Ca2NF
SrO Sr2NF Sr2NF
BaO Ba2NF Ba2NF

indicated the existence of both the MF2 and M2NF phases. All three systems had the
rock salt-type structure, similar to their respective MO oxide analog phases. The
unit cell parameters for Ca2NF, Sr2NF, and Ba2NF were reported as a = 4.937 Å,
5.368 Å, and 5.691 Å, respectively.
Pure calcium nitride fluoride phase was prepared from the reaction between cal-
cium nitride fluoride and CaF2 [91], and the temperature requirement is 900 °C for
24 h, whereas the oxynitride-fluoride phase was synthesized at 1000 °C, and the
reaction mixture is composed of CaO with Ca2NF. Sr2NF crystals were synthesized
by heating a melt of ratio 3:1 in the nickel crucible. Two crystals of dark red and
brownish yellow color are observed. These crystals are isostructural to Ca2NF
(Table 19.1).

19.5.2  Characteristics of TMNs

The basic synthetic approach used here mainly consists of ceramic methods to syn-
thesize the crystal of directed products from the melt. The techniques are summa-
rized as follows.
By using the standard ceramic technique, the ceramic mixture is heated to high
temperatures (1000–1500 °C) under the reaction atmospheres. The reaction used for
the preparation of the phase for CaMgNF is given below:

MgF2 + Mg + 2Ca + N → 2CaMgNF

The precursors for this reaction were available easily in the pure form. This
method is suitable for carrying out the reaction for a solid state. In order to melt the
reaction mixture, the reaction was heated till the mixture was melted. The melt pro-
cess can be used to synthesize nanocrystals. The reaction mixtures are heated to a
temperature so that all the contents start melting down. In case of extremely high
temperature, a flux is used which is composed of powder. These act as a solvent for
many reaction mixtures. The crystal growth demands gradual cooling to balance out
the growth and nucleation. Extremely slow nucleation dominates the formation of
large crystals, while the fast process directs the formation of powder. For the
synthesis-­related work, the starting material is moisture sensitive, and therefore they
are handled in an inert atmosphere. The inert atmosphere is achieved through the
use of glove bags [92–95]. The experimental setup for the synthesis of metal nitride
fluoride was carried out in nitrogen and argon through the furnace.
498 S. Kumar et al.

19.5.3  Thermal Stability

TGA technique is subjected to a controlled atmosphere to study the thermal proper-


ties. The mass of the sample is a temperature-based function. The instrument ele-
ment consists of a sample pan which is made from aluminum. This pan is further
located in the furnace and undergoes heating with the sample whenever the experi-
ment is carried out [96]. A thermocouple is placed within the furnace to regulate an
inert atmosphere. On completion of the experiment, the TGA curve appears on the
screen. Weight loss is observed from a descending curve [91]. The details such as
chemical composition, thermal stability, and reactivity could be inferred from
the curve.

19.6  Synthesis of Single-Metal Nitride Fluorides

19.6.1  Magnesium Nitride Fluoride System

In a glove bag filled with argon gas, the synthesized KCuF3 and KMgF3 compounds
were each used as precursors in separate reactions and mixed with magnesium
metal. The mixtures were each placed in separate nickel crucibles, which were posi-
tioned centrally in a nickel boat, and transferred into the Inconel tube from the glove
bag under argon gas [97]. The samples were then each heated under argon gas to
850 °C, followed by cooling to 200 °C. After that, the apparatus was reheated at
850 °C under nitrogen gas atmosphere. This temperature was maintained for 4 h
followed by cooling to 70 °C. The gas was terminated, and the crucible was removed
from the tube and placed in an argon-filled glove bag for observation. Below are the
chemical equations describing the attempted phase syntheses:

KCuF3 + 4 Mg + N 2 2Mg 2 NF + KF + Cu

KMgF3 + 3 Mg + N 2 2 Mg 2 NF + KF

19.6.2  Cobalt Nitride Fluoride System

Another ammonolysis reaction was carried out to make Co2NF. The proposed syn-
thesis of Co2NF started with the synthesis of NH4CoF3. The synthesized NH4CoF3
was put inside a nickel crucible. The treatment was done inside an argon-filled glove
bag due to the oxygen and moisture sensitivity of the reactant. The crucible with a
sample was inserted inside a molybdenum tube and then heated up to the decompo-
sition temperature of the precursor, 250 °C, under ammonia atmosphere for 2 h [98].
Then the system was cooled down to 70  °C.  Once the sample was cooled, it is
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 499

withdrawn for analysis after the termination of ammonia gas flow via X-ray powder
diffraction. This approach yielded a negative result, and NH4CoF3 was still intact
under this temperature.

19.7  Synthesis of Bimetallic Nitride Fluorides

Bimetallic nitride fluorides possess two metals combined with fluoride and nitride
ions. Such materials have not been widely studied.

19.7.1  Calcium Barium Nitride Fluoride

Another thermal treatment trial for attempting to make bimetallic nitride fluoride
was conducted. The mixture of two moles of calcium metal (Alfa Aesar 98.8%), one
mole barium metal (Alfa Aesar 99.2%) packaged in oil and one mole of barium
fluoride (Alfa Aesar 99.9%), was carried out in an evacuated glove bag to avoid
contamination of the reactants. Before mixing, excess oil was wiped from the chunk
of Ba using a polishing cloth, followed by placing the chunk in an Ar gas steam.
Following that, the mixture was loaded in a nickel crucible and then inserted into the
Inconel tube from the glove bag. After the insertion process, the tube was sealed
with a cap. Below is the chemical equation describing the targeted phase:

BaF2 + Ba + 2Ca ⇒ 2BaCaNF

Then the mixture was heated first under an argon flow followed by a nitrogen
flow using the following program (R = ramp function, L = temperature level, and
D = dwell time).

19.7.2  Calcium Magnesium Nitride Fluoride

The synthesis was based on thermal treatment of starting materials in the solid
phase. The mixture of two moles of Ca metal (Alfa Aesar 98.8%), one mole Mg
turnings (Alfa Aesar 99.98%) and one mole of MgF2 (Alfa Aesar 99.9%), was car-
ried out in an Ar-filled glove bag. These three compounds were weighed and loaded
in a nickel crucible and then inserted into the Inconel tube from the glove bag. The
chemical reaction for the proposed targeted compound is written below:

MgF2 + Mg + 2Ca + N 2 ⇒ 2CaMgNF


500 S. Kumar et al.

The mixture was heated first under an argon flow and followed by a nitrogen flow
using the following program (R  =  ramp function, L  =  temperature level, and
D = dwell time).

19.8  Multifunctional Hybrid-Structured Metal Nitrides

19.8.1  T
 itanium Nitride-Vanadium Nitride
(TiN-VN) Nanohybrids

Vanadium nitride (VN) has a greater limit in spite of low electronic conducting
properties [99]. Accordingly, consolidating VN and TiN into a productively quick
blended (electron and particle) transportation in the composites can be relied upon
to convey the elements for proficient charge transportation and electrochemical
vitality stockpiling. These materials have potential applications as super capacitive
devices. It has been shown that TiN-VN center shell-organized filaments are thought
to be a hopeful material for terminal supercapacitors.
Core/shell mesoporous hybrid-structured TiN-VN nanofibers were fabricated by
an electrospinning process. In this method, homogeneous solution was initially pre-
pared by mixing PVP polymer and Ti(OBu)4 in isopropyl liquor in room tempera-
ture for 12 h. To get ready vanadium pentoxide forerunner solution, PVP powder
mixed with a blend of ethylene glycol and ethanol, this was permitted to mix for
overnight at room temperature. In this way, for the addition of around 0.3  g of
vanadium(III) acetylacetonate in the arrangement, the final blend was mixed for
20 min period at 80 °C and cooled down to room temperature. Then, TiN-VN hybrid
core/shell nanofibers were fabricated under NH3 gas by heating to 800 °C for 1 h at
a pyrolysis ramp (25–800  °C). Cyclic voltammogramic (CV) experiments were
conducted for the evaluation of specific capacitance of these hybrid fibers. The
supercapacitor capacitance (C) was evaluated by the following equation [99]:

∫ I dcp
C=
2 mv∆V

where m is the electrode mass and I, v, ΔV, and j are the average charge and
discharge currents, scan rate, potential difference, and potential range, respectively.
The specific capacitance of the TiN-VN measured via galvanostatic charge-­
discharge experiments is calculated [100]:

I ∆t
C=
m∆V

Abbreviation C indicates the specific capacitance, I is the current, Δt is the time


interval, m is the mass of electrode, and Δv is a potential difference.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 501

19.8.2  Carbon Nitride/Ruthenium-Complex


Hybrid Photocatalysts

As a polymer semiconducting material carbon nitride (CN) was initially revealed in


1834 [83]. CN comprises two bounteous components present in earth and possesses
a band hole of around 2.8 eV. It is immiscible in acidic as well as base solutions and
additionally in different natural solvents. CN displays photocatalytic action for
water decrease and oxidation [15, 44]. CO2 is decreased by the photocatalytic action
of CN group [101]. The principle challenge includes figuring out how to stifle pro-
ton diminishment, which should all the more promptly (regarding thermodynamics
and energy) happen than when CO2 decreases.
Consolidated CN with ruthenium (Ru) complex is a notable impetus for the
reduction of CO2 [102]. High CO2 diminishment selectivity is shown under visible
light by crossover photocatalyst product. The crossbreed photocatalyst that we cre-
ated is equipped for changing over CO2 into formic acid (HCOOH) having over 80%
selectivity. It can be easily transportable and can be able to be promptly disinte-
grated into hydrogen (H2) and CO2 within the sight of the reasonable catalytic mate-
rial. We led isotopic tracing material that explores different avenues regarding CO2
and also some control analyses, the aftereffects of which bolster the response con-
spiracy appeared as shown in Fig. 19.2. From these analyses, primary objective facts
were mentioned. Initially, CN experiences excitation after noticeable light illumina-
tion. Second, the electrons in the conduction band of CN are exchanged toward most
minimally abandoned atomic orbital (LUMO) of adsorbed ruthenium buildings, and
oxidation of electron contributor species occurs at the valence band openings.

Fig. 19.2  IR-SEC of 4 in acetonitrile with Bu4NPF6 electrolyte. (4a) is formed by the first reduc-
tion of the complex to the radical anion with added electron density mostly centered on the bipyri-
dine ligand. (4b) follows through a ligand to metal charge transfer facilitated by the electron-donating
nature of the tert-butyl groups on the ligand. (4c) represents the second reduction of the complex
to the five-coordinate anion that acts as the active catalyst species. Reprinted with permission
from [102]
502 S. Kumar et al.

19.9  Applications of Hybrid-Structured Metal Nitrides

19.9.1  Hydrogen Generation with High Efficiency

One of the most straightforward sun-powered fuel is solar hydrogen. The routine of
agendas in light of sunlight-based hydrogen prompts a society that is steady, secure,
and natural. The current innovation is the production of sunlight-based fuel involv-
ing the participation of photovoltaic coordinate amalgamation of sun-oriented ener-
gies by simulated photosynthesis which has greater adaptability as far as framework
cost is concern.
Oxide photocatalysts which are extremely capable, for example, La-doped
NaTaO3 and Zn-doped Ga2O3, possess quantum efficiencies greater than 50% [103].
The photo-generated electrons at the conduction band of photocatalytic water frag-
ment develop hydrogen by reduction of water, whereas the gaps left in the valence
band cause water oxidation and release oxygen [104]. The competence of the con-
duction band has to be ought to be narrow compared to reversible hydrogen potential.

19.9.2  Photocatalytic Water Splitting

A strong arrangement between GaN and ZnO makes this photocatalyst most vigorous
(Ga1−xZnx)(N1−xOx), after it is modified with Rh-Cr-blended oxide co-catalyst [105].
The retention edge is at a wavelength of around 500 nm and quantum efficiency of
5.2% corresponding to light of wavelength of 410 nm. Diffuse reflectance spectra of
reflectance of (Ga1−xZnx)(N1−xOx) appeared in Fig. 19.3. GaN and ZnO are wide hole
semiconductors having wurtzite-type structure, with ingestion edges in shorter wave-
length than 400 nm (Ga1−xZnx)(N1−xOx) which has a more extended retention edge.
The description for the presence of the additional electronic state in the band hole was
explained by hypothetical computations and photoemission spectroscopy [106].
The silica particles were covered with a pearly glass plate along with blended
(Ga1−xZnx)(N1−xOx) photocatalyst. It has the most surprising solar-to-hydrogen
(STH) proficiency at display, yet the STH change effectiveness stays at around 0.2%
because it predominantly uses light photon with energy around 3 eV. The usage of
lower frequency light within sunlight-based illumination has the way for making
strides in proficiency.
Another auspicious possibility related to the photocatalytic water part is a
blended oxide having a perovskite structure. NaTaO3 and SrTiO3 consist of a
perovskite structure and demonstrate evident movement regarding general water-­
splitting reaction. NaTaO3 doped with La and altered with a NiO co-catalyst was
accounted for to possess half quantum proficiency for light of wavelength of
280 nm. In any case, these oxide photocatalytic materials which possess perovskite
structures consist of a retention edge in the UV locale and can’t use sunlight-based
illumination. Our system regarding the improvement of photocatalysts which
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 503

Fig. 19.3 UV-visible
diffuse reflectance spectra
for (Ga1−xZnx)(N1−xOx)
(x = 0.12): (a) as-prepared
and (b–g) after calcination
in air for 1 h at (b) 573, (c)
773, (d) 873, (e) 973, (f)
1073, and (g)
1173 K. Reprinted with
permission from [105]

possess the capacity to retain obvious light includes the substitution of oxygen
along with nitrogen. One of the oxygen particles of SrTiO3 can be replaced with a
nitrogen particle while adjusting the charge by substitution of Sr(II) with La(III),
yielding LaTiO2N [107]. Since the capability of the 2p orbital of nitrogen is thinner
than that of oxygen, the acquired oxynitrides have a smaller band hole than the
oxides. The diffuse reflectance spectra of LaTiO2N appears in Fig. 19.4. LaTiO2N
can retain light with a wavelength of 600–700 nm, and the benefit of LaTiO2N in the
assimilation spectra is self-evident [108].

19.9.3  P
 hotocatalytic Degradation Mechanism
of Environmental Pollutants

The mechanism of photocatalysis in g-C3N4/Ag/AgO ternary structured visible


light-driven photocatalyst is explained as follows [109] and shown in Fig. 19.5. The
photocatalytic action of g-C3N4 is upgraded after the adjustment by Ag and AgO,
over which is a fading photocatalytic productivity. Along these lines, to create har-
mony among the dynamic catching destinations is essential which hinder the recom-
bination of electron-opening sets, less caught parts prompting a bottom limit with
regard to the detachment of interfacial charge exchange. While illuminating the
Ag-stacked g-C3N4 photocatalyst, electrons in CB of g-C3N4 effectively get trans-
ferred to the metallic Ag via Schottky boundary in light of the fact that the CB of
g-C3N4 is superior to that of the stacked Ag/AgO, which is reliable for the past
examination of electron exchange between a semiconducting material and metal.
Electron exchange procedure occurs rapidly compared to electron-gap recombi-
nation in g-C3N4. A lot of CB electrons are put away in the Ag part. Hence, the
504 S. Kumar et al.

Fig. 19.4 Diffuse 2.5


reflection of UV-visible
spectra of 2.0
LaTiO2N. Reprinted with 1
permission from [107]
1.5

KM
1.0

0.5 2

0 300 400 500 600 700


Wavelength/nm

Fig. 19.5  Illustration of electron transfer in the ternary structured Ag/AgO/C3N4 catalyst.
Reprinted with permission from [109]

relocation, exchange, and recombination procedures of the photo-generated


electron-­opening sets inside a semiconductor.
For the preparation of g-C3N4/Ag hybrid photocatalysts using dimethylfor-
mamide (DMF) for reducing AgNO3 precursor, wet-chemical pathway is an effi-
cient approach. This catalyst showed superior photocatalytic performance under
visible light over bulk g-C3N4, and with an optimum Ag doping (~5  wt.%), the
photocatalytic efficiency reaches a significant kappa (κ) value of about 0.07 eV. The
improved catalytic performance is owing to the Ag particles, acting as a surface
plasmon to upsurge light absorption, and electron traps to stop the photo-generated
charge carrier recombination rate. This investigation is vital due to following two
reasons: (1) it is the first manifestation regarding the utility of dimethylformamide
(DMF) as an effective chemical-reducing reagent and polyvinylpyrrolidone (PVP)
for supporting the nano-phasic growth and uniform distribution of the plasmonic Ag
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 505

nanoparticles upon the surface of g-C3N4 nanosheets; and (2) it offers new advanced
material for its utilization in photocatalytic degradation of chemical dyes.

19.10  M
 etal Nitride-Based Hybrid Photocatalysts
for H2 Generation

The addition of nanomaterials that have a high electrocatalytic performance with a


semiconductor is established to be an effective method for solar energy conversion
[110]. Moreover, the efficient photocatalytic hydrogen generation can be achieved
by improving the reaction kinetics of the surface and transmission of interfacial
charge carriers [111]. Titanium nitride (TiN) has the ability as a co-catalyst to incor-
porate with semiconductors for efficient H2 evolution. TiO2/TiN core/shell nano-
belts have been synthesized by a hydrothermal method (by monitoring the
temperature and time) for H2 production without using any noble metal co-catalyst
[112]. A SEM and HR-TEM image of the prepared core/shell photocatalyst is
shown in Fig. 19.6.
The surface morphological image clearly has shown the nanobelt structure; in
[010] crystalline direction, TiO2 core nanobelts were developed with a lattice spac-
ing of 0.35 nm, and the shell formed with a lattice spacing of 0.21 nm of cubic TiN
as shown in Fig.  19.6b. The prepared core/shell photocatalyst showed stable and
high H2 generation value 120  μmol  h−1  g−1 compared to other catalysts. The
enhanced H2 generation is due to the improved efficiency of charge separation and
interfacial charge carrier’s transmission kinetics. Methanol is also likely to be used
as a new energy source, since it can be simply stored and transported. Hydrogen and
carbon monoxide (CO) can be produced by the decomposition of methanol. Nano-­
sized Ni metal deposited on hollow boron nitride spheres (HBNS) has been pre-
pared to generate the hydrogen from methanol [113]. The synthesized HBNS
showed a greater specific surface area of 229 m2/g with a 4-nm crystalline diameter.

Fig. 19.6 (a, b) SEM and HR-TEM images of TiO2/TiN core/shell nanobelts. Reprinted with
permission from [112]
506 S. Kumar et al.

By the decomposition of CH3OH, the prepared catalytic generates the H2 and exhib-
ited the extraordinary selectivity and activity with greater H2 conversion (>68%).
Similarly, boron nitride (BN) and titania-hydrogenated heterojunction of H-TiO2@
BN were prepared by using BN nanosheets as the photocatalyst support for H2 gen-
eration by He et al. [114]. The prepared hybrid catalyst significantly improved light
absorption ability, considerably reduced the recombination of charge carrier’s pair
and speed interfacial electron transfer, and reduced the resistance of the interfacial
charge transfer. As a result, the prepared H-TiO2@ BN catalyst exhibits the H2 pro-
duction of 721 μmol.
A novel compound of Ni nanoparticles supported boron nitride spheres, and
sheets have been prepared by an impregnation chemical reduction method for
hydrogen evolution with different Ni atomic contents of 6.8, 9.0, 9.2, and 12 wt.%
[115]. The as-prepared Ni-supported boron nitride catalysts presented a greater
catalytic activity for H2 production compared with the unsupported Ni catalyst.
9.0 wt.% Ni-supported sample exhibits the higher H2 generation rate of 476 mL.
min−1.g−1 and showed good recycling performance. Similarly, highly dispersed cop-
per nanoparticles on hexagonal boron nitride (h-BN) nanosheets have been prepared
by a facile solution-phase synthesis for H2 generation from ammonia borane (AB)
[116]. Due to the high distribution of Cu nanoparticles on h-BN, the composite
shows greater hydrogen generation through the hydrolysis of AB.  Moreover, at
room temperature, the catalytic hydrogen generation of Cu/h-BN is dependent on
the copper content. The 29.4 wt.% Cu/h-BN composite exhibited the best catalytic
H2 generation. 29.4  wt.% Cu/h-BN composite sample showed lesser activation
energy (23.8  kJ  mol−1) for the reaction. Therefore, 29.4  wt.% Cu/h-BN sample
exhibits greater H2 generation of 86 mL min−1 as compared to other samples due to
the suitable wt.% of Cu content. These composites progress a catalytic activity for
catalytic hydrogen production from aqueous AB.
Arunachalam et al. reported titanium oxynitride (TiOxNy) nanoparticle grown on
TiO2 nanorod (TNR) arrays using a facile nitridation hydrothermal technique [117].
The prepared photocatalyst displays the high photocurrent density (J) 2.1 mA/cm2
at 1.23 V in contrast with a pure sample. The TiOxNy-decorated TNR arrays exhib-
ited threefold improvement which was achieved due to the extraordinary absorption
of visible light and speed separation of charge carrier. Furthermore, the addition of
TiOxNy layer on the TNR surface decreases the resistance of the interfacial in the
solid-liquid region, and the TiOxNy layer reduced the surface reactions where the
charge recombination reaction often occurs, which leads to the enhancement of
photoelectrochemical activity.
Gallium nitride (GaN) is one of the greatest capable materials for photoelectrode
owing to its tunable band-edge potentials and has an extraordinary stability due to
its strong GaN ionic bonding, superior light absorption ability, and amazing proper-
ties of charge carrier transportation. Moreover, its both minimum conduction band
and maximum valence band values are appropriate for water splitting. GaN nanow-
ires have been grown on sapphire (Al2O3) and silicon (Si) substrate using a chemical
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 507

Fig. 19.7 (a) SEM image and (b) H2 generation of GaN nanowires. Reprinted with permission
from [118]

vapor deposition method for water splitting [118]. The SEM image of prepared GaN
nanowires is shown in Fig. 19.7a. The crystal structure analysis confirmed that the
prepared GaN nanowires are of great crystalline quality with hexagonal wurtzite
structure.
The prepared GaN/Si nanowires display substantial improvement in the move-
ment of photo-excited hole to the interface (semiconductor-electrolyte) when com-
pared with the GaN/Al2O3 nanowires. Moreover, GaN/Si nanowires display the
significant improvement of photocurrent density (0.15  mA/cm2) compared with
GaN/Al2O3 nanowires. At the visible region, GaN/Si nanowires also presented a
substantial improvement in H2 generation, which is credited to its heterostructure,
reduced photo-generated charge carriers, and prolonged the responses of visible
light. Multilayered ITO/V-doped TiO2 has been fabricated using a radiofrequency
magnetron sputtering for efficient H2 generation [119]. Owing to the impurity
energy levels created in the bandgap of V-doped TiO2, it shows improved optical
absorption efficiency. But, due to the increased photo-excited charge carrier recom-
bination rate, the small photocurrent is observed with a single-bilayer ITO/V-doped
TiO2 sample. To decrease the charge carrier recombination rate, a multilayer ITO/V-­-
doped TiO2 sample has been prepared. The multilayer sample considerably dimin-
ishes the recombination rate due to the produced photoelectrons moved into a
limited thickness of TiO2 film, then quickly enters the interface, and then is imme-
diately injected into the ITO layer. The maximum photocurrent value is attained
with multilayer sample. The improved photocurrent is accredited to higher visible
light absorption and the space-charge interfaces. As a result, the higher H2 produc-
tion rate of 31.2 μ mol/h is obtained compared to single-layer sample as shown in
Fig.  19.7b. Similarly, several researchers have reported the photoelectrochemical
properties of metal nitrides for water-splitting applications and photoredox catalysis
[120–122].
508 S. Kumar et al.

19.10.1  g-C3N4-Based Hybrid Photocatalysts for H2 Generation

In recent years, graphitic carbon nitride (g-C3N4) has been attracted as an emerging
photocatalyst due to its distinctive structure, suitable properties for water splitting,
high visible light absorption ability, earth rich, and better thermal and chemical
stability. However, due to its lesser surface area and higher recombination rate of
photo-generated charge carriers display obstructing its photocatalytic efficiency
[123]. Therefore, the formation of heterojunctions by the combinations of a semi-
conductor composite is one of the greatest techniques to attaining effective photo-
catalyst properties. Due to the interface actions between the heterojunctions
resulting in enhances the usage of visible light, speed separation, and migration of
charge and more effective creation of the oxidizing species [124]. α-FeOOH acting
as an outstanding OER co-catalyst supports for the photoelectrochemical water-­
splitting activity owing to its structures significantly enabling the electrolyte trans-
port properties. Hence, Luo et  al. [125] reported the Z-scheme visible light
heterostructure. β-FeOOH/g-C3N4 composite has been prepared for H2 generation.
The synthesis procedure of the heterostructure composite flow chart is shown in
Fig. 19.8.
Here, the modified g-C3N4 with β-FeOOH composite photocatalyst showed a
higher H2 generation rate of 2.02 mmol·h−1·g−1 (six times higher) compared to that
of the pure sample. The noticeably enhanced photocatalytic H2 generation was cred-
ited to the significantly improved charge carrier separation efficiency by creating
spatial separated photo-generated charge carriers in the Z-scheme heterojunction,
corresponding to the conduction and the valence band of g-C3N4 and β-FeOOH,
respectively, therefore stopping the charge carrier recombination rate. The suitable
heterojunction interfacial interaction shows a critical role in interfacial charge car-
rier’s separation/movement. The novel strategy of two-dimensional surface-to-­
surface heterojunction is an active technique for enhancing the catalytic performance
since better contact area could improve the transfer rate of interfacial charge.
Ultrathin S-scheme heterojunction WO3/g-C3N4 2D/2D photocatalyst has been syn-
thesized using an electrostatic self-assembly technique with a thickness of
2.5–3.5 nm for efficient H2 generation [126].
Figure 19.9a shows the schematic representation of WO3/g-C3N4 2D/2D hetero-
junctions. Figure 19.9b displays the H2 production comparison rates of all samples.
Owing to the low conduction position, WO3 nanosheets showed no H2 production.

Fig. 19.8  CN/Fe-x composite preparation flow chart. Reprinted with permission from [125]
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 509

Fig. 19.9 (a) Formation diagram of 2D/2D WO3/g-C3N4 heterojunctions, (b) H2 production (the
symbols 10%, 15%, 20%, and 30% on the x-axis represent the 10%, 15%, 20%, and 30% of WO3),
and (c–e) the work functions of g-C3N4 and WO3 before contact, after contact, and the S-scheme
charge transfer mechanism under light irradiation. Reprinted with permission from [126]

Owing to higher conduction band position and photo-excited electrons with suffi-
cient ability of reduction, pure g-C3N4 nanosheets exhibited H2 production. But,
with the occurrence of WO3 nanosheets, the WO3/g-C3N4 composite showed
enhanced H2 generation. Particularly, the 15%WO3/g-C3N4 composite sample shows
the maximum H2 generation activity (982  μmol/h/g), which is about 1.7 times
greater than that of bare g-C3N4 sample. Figure 19.9c–e shows the S-scheme mecha-
nism of charge transfer between WO3 and g-C3N4. Generally, g-C3N4 has a higher
Fermi level with a smaller work function (4.18 eV) reduction-type catalyst.
In contrast, WO3 has a lower Fermi level with greater work function (6.23 eV)
oxidation-type catalyst (Fig. 19.9c). When these two materials are in close interac-
tion, g-C3N4 electrons moved to WO3 through the interface until their Fermi levels
are the same (Fig. 19.9d). Upon light illumination (Fig. 19.9e), the mechanism of
S-scheme heterojunction will remove the comparatively useless charge carriers but
keep beneficial charge carriers. Hence, this charge carrier movement procedure pro-
vides the composite with the highest redox capability, therefore giving a strong
moving force for running catalytic reaction.
Similarly, 1D graphitic carbon nitride/graphene/recycled carbon fiber (g-CN/G/
RCF) heterostructure composite has been prepared by using 1D RCF and facile
steam activation scheme for H2 generation [127]. The formation of g-CN/G/RCF is
shown in Fig. 19.10a. The combination of melamine/graphene/RCF was heated at
550 °C for 3 h to prepare the g-CN/G/RCF30 composite. The SEM and TEM images
of optimized sample g-CN/G/RCF30-10 are shown in Fig.  19.10b, c. Due to the
outstanding electric conductivity and strong attraction of graphene to g-C3N4, the
as-prepared g-CN/G/RCF composites presented improved catalytic H2 generation.
The optimized g-CN/G/RCF sample exhibited the uppermost catalytic H2 genera-
tion rate of 411.6 μmol h−1 g−1, which is about 3.5, 2.1 times greater than those of
pure g-C3N4 and g-CN/RCF (absence of graphene). Moreover, is showed outstand-
ing photo-stability for H2 production, showing no substantial loss on H2 generation
rate within 24 h of the catalytic reaction (Table 19.2).
510 S. Kumar et al.

Fig. 19.10 (a) Formation of g-CN/G/RCFt and (b, c) SEM and TEM images for g-CN/G/RCF30-10
sample. Reprinted with permission from [127]

Table 19.2  Graphitic carbon nitride (g-C3N4)-based hybrid photocatalysts for H2 generation
H2
g-C3N4-based hybrid Reaction solution/ production
photocatalysts electrolyte Light source μmol h−1 g−1 Refs.
Zn porphyrin/C3N4 C6H8O6 Monochromatic light 524 [128]
(λ ≥ 420 nm)
g-C3Nx C6H15NO3 Xe lamp 403.1 [129]
(λ ≥ 420 nm)
Pt/ITO/g-C3N4 C6H15NO3 Xe arc lamp 335.5 [130]
Pd/2D-C3N4 C6H15NO3 Xe lamp 1208.6 [131]
(λ ≥ 420 nm)
2D-Ni2P@BP/CN C6H15NO3 Xe lamp 858.2 [132]
(λ ≥ 420 nm)
K-doped g-C3N4 C6H15NO3 Xe lamp 919.5 [133]
(λ ≥ 420 nm)
Porous g-C3N4 C6H15NO3 Xe lamp 90 [134]
(λ ≥ 420 nm)
S-doped g-C3N4 C6H15NO3 Xe lamp 1511.2 [135]
(λ ≥ 420 nm)
MoS2/g-C3N4 C6H15NO3 Xe lamp 1155 [136]
(λ ≥ 420 nm)
g-C3N4 C6H15NO3 Xe lamp 30.67 [137]
(λ ≥ 420 nm)
(continued)
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 511

Table 19.2 (continued)
H2
g-C3N4-based hybrid Reaction solution/ production
photocatalysts electrolyte Light source μmol h−1 g−1 Refs.
CoS2/g-C3N4 C6H15NO3 Visible light 1232 [138]
(λ > 420 nm
PtNi/g-C3N4 C6H15NO3 Visible light 104.7 [139]
(λ > 420 nm
MoS2/g-C3N4 H2SO4 Visible light 867.6 [140]
(λ > 420 nm
PtAu/g-C3N4 Na2SO3/Na2S Xe lamp 1009 [141]
(λ ≥ 420 nm)
Metakaolin/g-C3N4 C6H15NO3 Xe lamp 288 [142]
(λ ≥ 420 nm)

19.11  Conclusions

In summary, we have reviewed the recent developments in different metal nitrides


and graphitic carbon nitrides to achieve efficient H2 generation. Our discussions
include different synthesize approaches of different metal nitrides, crystal struc-
tures, as well as efficient hydrogen evaluation. In general, absorption of light, sepa-
ration of charge carrier’s pair, passage of hole, surface kinetics, suitable structure,
and optimized composition of materials are the important features in research and
design. Different metal nitrides and g-C3N4-based hybrid composite systems for
catalytic H2 generation have been studied. Metal nitrides are the most appropriate
photocatalysts that can be used in various photocatalysis methods. A combination of
nano-sized metal nitrides with appropriate small bandgap materials produces hybrid
catalysts which show development in H2 generation from the visible light region.
Still, advanced technological developments like doping and ease growth mecha-
nisms are being subtle. It is also intended in this book chapter that understanding of
other material and compositional systems is necessary to join the whole solar spec-
trum. However, still research works are going on to integrate metal nitride nanow-
ires with other materials which still need improvement toward attaining efficient H2
generation to meet the future needs. In conclusion, the following upcoming research
aspects of metal nitrides for their widespread usage in practical applications include
defect-free crystallite structures, desired thickness to prevent misfit disorders, con-
trolled doping, enhanced surface area, less cost, and consistency. Moreover, the
progress of new developments with a strong theoretical background is necessary for
a better sympathetic mechanism for hydrogen generation and eco-friendly water-­
splitting procedure for H2 generation.

Acknowledgments  This work was supported by the National Research Foundation of Korea
grant funded by the Korea government (No. NRF-2017R1A4A1015581).
512 S. Kumar et al.

References

1. Yerga, R. M. N., Galvan, M. C. A., del Valle, F., de la Mano, J. A. V., & Fierro, J. L. G. (2009).
Water splitting on semiconductor catalysts under visible-light irradiation. ChemSusChem, 2,
471–485.
2. Herrmann, J. M. (1999). Heterogeneous photocatalysis: Fundamentals and applications to the
removal of various types of aqueous pollutants. Catalysis Today, 53, 115–129.
3. Lützenkirchen-Hecht, D., & Frahm, R. (2005). Structure of reactively sputter deposited tin-­
nitride thin films: A combined X-ray photoelectron spectroscopy, in situ X-ray reflectivity
and X-ray absorption spectroscopy study. Thin Solid Films, 493, 67–76.
4. Qu, F., Yuan, Y., & Yang, M. (2017). Designed synthesis of Sn3N4 nanoparticles through soft
urea route with excellent gas sensing properties. Chemistry of Materials, 29(3), 969–974.
5. Li, X., Hector, A. L., Owen, J. R., & Shah, S. I. U. (2016). Evaluation of nanocrystalline
Sn3N4 derived from ammonolysis of Sn (NEt2)4 as a negative electrode material for Li-ion
and Na-ion batteries. Journal of Materials Chemistry A, 4, 5081–5087.
6. Yu, X., & Pickup, P. G. (2008). Recent advances in direct formic acid fuel cells (DFAFC).
Journal of Power Sources, 182, 124–132.
7. Jin, J., He, Z., & Zhao, X. (2020). Effect of Al content on the corrosion resistance and conduc-
tivity of metal nitride coating in the cathode environment of PEMFCs. Materials Chemistry
and Physics, 245, 122739.
8. Kojima, Y., Kawai, Y., & Ohba, N. (2006). Hydrogen storage of metal nitrides by a mechano-
chemical reaction. Journal of Power Sources, 159, 81–87.
9. Takai, O. (1987). A new electrochromic system using tinnitride thin-films. Proceedings of the
SID, 10014, 243–246.
10. Inoue, Y., Nomiya, M., & Takai, O. (1998). Physical properties of reactive sputtered tin-­
nitride thin films. Vacuum, 51, 673–676.
11. Chatterjee, D., & Dasgupta, S. (2005). Visible light induced photocatalytic degradation of
organic pollutants. Journal of Photochemistry and Photobiology C, 6, 186–205.
12. Fujishima, A., & Honda, K. (1972). Electrochemical photolysis of water at a semiconductor
electrode. Nature, 238, 37.
13. Radoičić, M. B., Janković, I. A., Despotović, V. N., Šojić, D. V., Savić, T. D., Šaponjić, Z. V.,
Abramović, B. F., & Čomor, M. I. (2013). The role of surface defect sites of titania nanopar-
ticles in the photocatalysis: Aging and modification. Applied Catalysis B: Environmental,
138–139, 122–127.
14. Li, T., He, Y., Lin, H., Cai, J., Dong, L., Wang, X., Luo, M., Zhao, L., Yi, X., & Weng,
W. (2013). Synthesis, characterization and photocatalytic activity of visible-light plasmonic
photocatalyst AgBr-SmVO4. Applied Catalysis B: Environmental, 138–139, 95–103.
15. Wang, X., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J. M., Domen, K., &
Antonietti, M. (2008). A metal-free polymeric photocatalyst for hydrogen production from
water under visible light. Nature Materials, 8, 76.
16. Zheng, Y., Li, X., Pi, C., Song, H., Gao, B., Chu, P. K., & Huo, K. (2020). Recent advances
of two-dimensional transition metal nitrides for energy storage and conversion applications.
FlatChem, 19, 100149.
17. Cheng, L., Xie, S., Zou, Y., Ma, D., Sun, D., Li, Z., Wang, Z., & Shi, J.-W. (2019). Noble-­
metal-­free Fe2P–Co2P co-catalyst boosting visible-light-driven photocatalytic hydrogen pro-
duction over graphitic carbon nitride: The synergistic effects between the metal phosphides.
International Journal of Hydrogen Energy, 44, 4133–4142.
18. Li, H., Hu, H., Bai, C., Bao, C., Feng, Z., & Guo, F. (2019). The metal-free magnetism and
ferromagnetic narrow gap semiconductor properties in graphene-like carbon nitride. Physica
B, 555, 91–95.
19. Wu, P., Lu, L., He, J., Chen, L., Chao, Y., He, M., Zhu, F., Chu, X., Li, H., Zhu, W., et al.
(2020). Green Energy Environment, 5, 166–172.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 513

20. Ji, D., Chong, X. Y., & Feng, J. (2019). Electronic, mechanical and hydrogen storage proper-
ties of novel Mg3N2. Journal of Alloys and Compounds, 800, 8–15.
21. Chalé-Lara, F., Farías, M. H., De la Cruz, W., & Zapata-Torres, M. (2010). Influence of depo-
sition temperature on the structural and morphological properties of Be3N2 thin films grown
by reactive laser ablation. Applied Surface Science, 256, 7628–7631.
22. Adhikari, V., Liu, Z.  T. Y., Szymanski, N.  J., Khatri, I., Gall, D., Sarin, P., & Khare,
S. V. (2018). First-principles study of mechanical and magnetic properties of transition metal
(M) nitrides in the cubic M4N structure. Journal of Physics and Chemistry of Solids, 120,
197–206.
23. Shi, F., Wang, L. J., Cui, W. F., & Liu, C. M. (2007). Precipitation behavior of M2N in a
high-nitrogen austenitic stainless steel during isothermal aging. Acta Metallurgica Sinica,
20, 95–101.
24. Chen, Y., Liu, X., Hou, L., Guo, X., Fu, R., & Sun, J. (2020). Construction of covalent
bonding oxygen-doped carbon nitride/graphitic carbon nitride Z-scheme heterojunction for
enhanced visible-light-driven H2 evolution. Chemical Engineering Journal, 383, 123132.
25. Cheng, H., Zhao, K., Gong, Y., Wang, X., Wang, R., Wang, F., Hu, R., Wang, F., Zhang, X.,
He, J., & Tian, X. (2020). Covalent coupling regulated thermal conductivity of poly(vinyl
alcohol)/boron nitride composite film based on silane molecular structure. Composites Part
B: Engineering, 137, 106026.
26. Takagaki, A. (2020). Effects of post-thermal treatments of ball-milled boron nitrides on solid
base catalysis. Catalysis Today, 352, 279–286.
27. Dhanumalayan, E., Joshi, G. M., Kaleemulla, S., Deshmukh, R. R., & Kumar, S. M. S. (2019).
Physico-chemical and surface properties of air plasma treated PVDF/PMMA/Attapulgite/
hexagonal-Boron Nitride blends. Progress in Organic Coatings, 131, 17–26.
28. Mallick, S., & Kumar, P. (2020). Computational evidence for sulfur atom tunneling in the
ring flipping reaction of S4N4. Chemical Physics Letters, 749, 137440.
29. Kelly, P. F., & Woollins, J. D. (1986). The preparation and structure of complexes containing
simple sulphur-nitrogen ligands. Polyhedron, 5, 607–632.
30. Su, F., Antonietti, M., & Wang, X. (2012). Mpg-C3N4 as a solid base catalyst for Knoevenagel
condensations and transesterification reactions. Catalysis Science and Technology, 2,
1005–1009.
31. Wang, X. C., Maeda, K., Homas, T. A., Takanabe, K., Xin, G., Carlsson, J. M., Domen, K., &
Antonietti, M. A. (2009). Metal-free polymeric photocatalyst for hydrogen production from
water under visible light. Nature Materials, 8, 76–80.
32. Zhang, L., Wang, H., Shen, W., Qin, Z., Wang, J., & Fan, W. (2016). Controlled synthesis of
graphitic carbon nitride and its catalytic properties in Knoevenagel condensations. Journal of
Catalysis, 344, 293–302.
33. Alexandrescu, R., Huisken, F., Pugna, G., Crunteanu, A., Petcu, S., Cojocaru, S., Cireasa, R.,
& Morjan, I. (1997). Preparation of carbon nitride fine powder by laser induced gas-phase
reactions. Applied Physics A: Materials Science & Processing, 65, 207–213.
34. Yin, L. W., Bando, Y., Li, M. S., Liu, Y. X., & Qi, Y. X. (2003). Unique single-crystalline beta
carbon nitride nanorods. Advanced Materials, 15(21), 1840–1844.
35. Jun, Y.-S., Park, J., Lee, S. U., Thomas, A., Hong, W. H., & Stucky, G. D. (2013). Three-­
dimensional macroscopic assemblies of low dimensional carbon nitrides for enhanced hydro-
gen evolution. Angewandte Chemie International Edition, 52, 11083–11087.
36. Shang, L., Bian, T., Zhang, B., Zhang, D., Wu, L.-Z., Tung, C.-H., Yin, Y., & Zhang, T. (2014).
Graphene-supported ultrafine metal nanoparticles encapsulated by mesoporous silica: Robust
catalysts for oxidation and reduction reactions. Angewandte Chemie International Edition,
53, 250–254.
37. Zhang, X., Wang, H., Wang, H., Zhang, Q., Xie, J., Tian, Y., Wang, J., & Xie, Y. (2014).
Single-layered graphitic-C3N4 quantum dots for two-photon fluorescence imaging of cellular
nucleus. Advanced Materials, 26, 4438–4443.
514 S. Kumar et al.

38. Yang, J., Zhang, H., Chen, B., Tang, H., Li, C., & Zhang, Z. (2015). Fabrication of the
g-C3N4/Cu nanocomposite and its potential for lubrication applications. RSC Advances, 5,
64254–64260.
39. Ong, W.-J., Tan, L.-L., Ng, Y. H., Yong, S.-T., & Chai, S.-P. (2016). Graphitic carbon nitride
(g C3N4) based photocatalysts for artificial photosynthesis and environmental remediation:
Are we a step closer to achieving sustainability? Chemical Reviews, 116, 7159–7329.
40. Suresh, S., & Irvine, A. E. (2015). The NOTCH signaling pathway in normal and malignant
blood cell production. Journal of Cell Communication and Signaling, 9(1), 5–13.
41. Corma, A., Atienzar, P., García, H., & Chane-Ching, J.-Y. (2004). Hierarchically mesostruc-
tured doped CeO2 with potential for solar-cell use. Nature Materials, 3, 394.
42. Morshed, A. H., Moussa, M. E., Bedair, S. M., Leonard, R., Liu, S. X., & El-Masry, N. (1997).
Violet/blue emission from epitaxial cerium oxide films on silicon substrates. Applied Physics
Letters, 70, 1647–1649.
43. Tsunekawa, S., Fukuda, T., & Kasuya, A. (2000). Blue shift in ultraviolet absorption spectra
of monodisperse CeO2−x nanoparticles. Journal of Applied Physics, 87, 1318–1321.
44. Dunkle, S. S., Helmich, R. J., & Suslick, K. S. (2009). BiVO4 as a visible-light photocatalyst
prepared by ultrasonic spray pyrolysis. Journal of Physical Chemistry C, 113, 11980–11983.
45. Wang, Y., Wang, X., & Antonietti, M. (2012). Polymeric graphitic carbon nitride as a het-
erogeneous organocatalyst: From photochemistry to multipurpose catalysis to sustainable
chemistry. Angewandte Chemie. International Edition, 51, 68–89.
46. Su, F., Mathew, S.  C., Möhlmann, L., Antonietti, M., Wang, X., & Blechert, S. (2011).
Aerobic oxidative coupling of amines by carbon nitride photocatalysis with visible light.
Angewandte Chemie. International Edition, 50, 657–660.
47. Guo, Y., Chu, S., Yan, S., Wang, Y., & Zou, Z. (2010). Developing a polymeric semiconductor
photocatalyst with visible light response. Chemical Communications, 46, 7325–7327.
48. Chen, X., Jun, Y.-S., Takanabe, K., Maeda, K., Domen, K., Fu, X., Antonietti, M., & Wang,
X. (2009). Ordered mesoporous SBA-15 type graphitic carbon nitride: A semiconductor host
structure for photocatalytic hydrogen evolution with visible light. Chemistry of Materials, 2,
4093–4095.
49. Li, X.-H., Zhang, J., Chen, X., Fischer, A., Thomas, A., Antonietti, M., & Wang, X. (2011).
Condensed graphitic carbon nitride nanorods by nanoconfinement: Promotion of crystallinity
on photocatalytic conversion. Chemistry of Materials, 23, 4344–4348.
50. Chen, X., Zhang, J., Fu, X., Antonietti, M., & Wang, X. (2009). Fe-g-C3N4-catalyzed oxida-
tion of benzene to phenol using hydrogen peroxide and visible light. Journal of the American
Chemical Society, 131, 11658–11659.
51. Wang, X., Chen, X., Thomas, A., Fu, X., & Antonietti, M. (2009). Metal-containing carbon
nitride compounds: A new functional organic–metal hybrid material. Advanced Materials,
21, 1609–1612.
52. Ge, L., Han, C., Liu, J., & Li, Y. (2011). Enhanced visible light photocatalytic activity of
novel polymeric g-C3N4 loaded with Ag nanoparticles. Applied Catalysis A: General,
409–410, 215–222.
53. Wang, X., Maeda, K., Chen, X., Takanabe, K., Domen, K., Hou, Y., Fu, X., & Antonietti,
M. (2009). Polymer semiconductors for artificial photosynthesis: Hydrogen evolution by
mesoporous graphitic carbon nitride with visible light. Journal of the American Chemical
Society, 131, 1680–1681.
54. Gust, D., Moore, T. A., & Moore, A. L. (2001). Mimicking photosynthetic solar energy trans-
duction. Accounts of Chemical Research, 34, 40–48.
55. Lonergan, M.  C., Cheng, C.  H., Langsdorf, B.  L., & Zhou, X. (2002). Electrochemical
characterization of polyacetylene ionomers and polyelectrolyte-mediated electrochemistry
toward interfaces between dissimilarly doped conjugated polymers. Journal of the American
Chemical Society, 124, 690–701.
56. Wong, W. Y., Wang, X. Z., He, Z., Djurišić, A. B., Yip, C. T., Cheung, K. Y., Wang, H., Mak,
C. S. K., & Chan, W. K. (2007). Metallated conjugated polymers as a new avenue towards
high-efficiency polymer solar cells. Nature Materials, 6, 521.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 515

57. Zhang, J., Chen, X., Takanabe, K., Maeda, K., Domen, K., Epping, J. D., Fu, X., Antonietti,
M., & Wang, X. (2010). Synthesis of a carbon nitride structure for visible-light catalysis by
copolymerization. Angewandte Chemie International Edition, 49, 441–444.
58. Ha, Y., Choi, M.-C., Kim, I., Ha, C.-S., Kim, Y., & Han, M. (2010). Microstructure and prop-
erties of rigid rod-like polyimide/flexible coil-like poly(amide-imide) molecular composite
films. Macromolecular Research, 18, 14–21.
59. Yan, S. C., Li, Z. S., & Zou, Z. G. (2009). Photodegradation performance of g-C3N4 fabri-
cated by directly heating melamine. Langmuir, 25, 10397–10401.
60. Vyas, V. S., Haase, F., Stegbauer, L., Savasci, G., Podjaski, F., Ochsenfeld, C., & Lotsch,
B. V. (2015). A tunable azine covalent organic framework platform for visible light-induced
hydrogen generation. Nature Communications, 6, 8508.
61. Li, P., Wang, M., Cheng, J., Xu, C., & Lu, H. (2013). Spectral hashing with semantically
consistent graph for image indexing. IEEE Transactions on Multimedia, 15, 141–152.
62. Ali, I., Park, S., & Kim, J.-O. (2020). Modeling the photocatalytic reactions of g-C3N4-TiO2
nanocomposites in a recirculating semi-batch reactor. Journal of Alloys and Compounds,
821, 153498.
63. Yang, X., Tian, Z., Chen, Y., Huang, H., & Hu, J. (2020). One-pot calcination preparation
of graphene/g–C3N4–Co photocatalysts with enhanced visible light photocatalytic activity.
International Journal of Hydrogen Energy, 45, 12889–12902.
64. Li, T., Zhao, L., He, Y., Cai, J., Luo, M., & Lin, J. (2013). Synthesis of g-C3N4/SmVO4 com-
posite photocatalyst with improved visible light photocatalytic activities in RhB degradation.
Applied Catalysis B: Environmental, 129, 255–263.
65. Yan, S. C., Lv, S. B., Li, Z. S., & Zou, Z. G. (2010). Organic–inorganic composite photo-
catalyst of g-C3N4 and TaON with improved visible light photocatalytic activities. Dalton
Transactions, 39, 1488–1491.
66. Li, W., Li, C., Chen, B., Jiao, X., & Chen, D. (2015). Facile synthesis of sheet-like N–TiO2/g-­
C3N4 heterojunctions with highly enhanced and stable visible-light photocatalytic activities.
RSC Advances, 5, 34281–34291.
67. Yan, X., Xu, R., Guo, J., Cai, X., Chen, D., Huang, L., Xiong, Y., & Tan, S. (2017). Enhanced
photocatalytic activity of Cu2O/g-C3N4 heterojunction coupled with reduced graphene oxide
three-dimensional aerogel photocatalysis. Materials Research Bulletin, 96, 18–27.
68. Liu, Y., Chen, P., Chen, Y., Lu, H., Wang, J., Yang, Z., Lu, Z., Lia, M., & Fangc, L. (2016). In
situ ion-exchange synthesis of SnS2/g-C3N4 nanosheets heterojunction for enhancing photo-
catalytic activity. RSC Advances, 6, 10802–10809.
69. Flaherty, D. W., May, R. A., Berglund, S. P., Stevenson, K. J., & Mullins, C. B. (2010). Low
temperature synthesis and characterization of nanocrystalline titanium carbide with tunable
porous architectures. Chemistry of Materials, 22, 319–329.
70. Lengauer, W. (2008). Transition metal carbides, nitrides, and carbonitrides. In Hand book of
ceramic hard materials (pp. 202–252). Wiley-VCH Verlag GmbH.
71. Penrice, T. W. (2000). Tungsten compounds. Kirk-Othmer encyclopedia of chemical technol-
ogy. Wiley.
72. Fu, Z.-W., Wang, Y., Yue, X.-L., Zhao, S.-L., & Qin, Q.-Z. (2004). Electrochemical reactions
of lithium with transition metal nitride electrodes. The Journal of Physical Chemistry. B, 108,
2236–2244.
73. Oyama, S. T. (1996). Introduction to the chemistry of transition metal carbides and nitrides.
In S. T. Oyama (Ed.), The chemistry of transition metal carbides and nitrides (pp. 1–27).
Dordrecht: Springer Netherlands.
74. Sun, Q., & Fu, Z.-W. (2008). Vanadium nitride as a novel thin film anode material for
rechargeable lithium batteries. Electrochimica Acta, 54, 403–409.
75. Choi, D., Blomgren, G. E., & Kumta, P. N. (2006). Fast and reversible surface redox reaction
in nanocrystalline vanadium nitride supercapacitors. Advanced Materials, 18, 1178–1182.
76. Bi, W., Hu, Z., Li, X., Wu, C., Wu, J., Wu, Y., Wu, Y., & Xie, Y. (2015). Metallic mesocrystal
nanosheets of vanadium nitride for high-performance all-solid-state pseudocapacitors. Nano
Research, 8, 193–200.
516 S. Kumar et al.

77. Dirks, A. G., Wolters, R. A. M., & De Veirman, A. E. M. (1992). Columnar microstructures
in magnetron-sputtered refractory metal thin films of tungsten, molybdenum and W-Ti-(N).
Thin Solid Films, 208, 181–188.
78. Dong, S., Chen, X., Gu, L., Zhou, X., Li, L., Liu, Z., Han, P., Xu, H., Yao, J., Wang, H.,
Zhang, X., Shang, C., Cui, G., & Chen, L. (2011). One dimensional MnO2/titanium nitride
nanotube coaxial arrays for high performance electrochemical capacitive energy storage.
Energy & Environmental Science, 4, 3502–3508.
79. Li, G.-r., Wang, F., Jiang, Q.-w., Gao, X.-p., & Shen, P.-w. (2010). Carbon nanotubes with
titanium nitride as a low-cost counter-electrode material for dye-sensitized solar cells.
Angewandte Chemie International Edition, 49, 3653–3656.
80. Zhu, C., Yang, P., Chao, D., Wang, X., Zhang, X., Chen, S., Tay, B. K., Huang, H., Zhang,
H., Mai, W., & Fan, H. J. (2015). All metal nitrides solid-state asymmetric supercapacitors.
Advanced Materials, 27, 4566–4571.
81. Becker, J. S., Suh, S., Wang, S., & Gordon, R. G. (2003). Highly conformal thin films of
tungsten nitride prepared by atomic layer deposition from a novel precursor. Chemistry of
Materials, 15, 2969–2976.
82. Rische, D., Parala, H., Gemel, E., Winter, M., & Fischer, R. A. (2006). New tungsten(VI)
guanidinato complexes: Synthesis, characterization, and application in metal−organic chemi-
cal vapor deposition of tungsten nitride thin films. Chemistry of Materials, 18, 6075–6082.
83. Liebig, J. (1834). Uber einige Stickstoff—Verbindungen. Annalen der Pharmacie,
10(1), 1–47.
84. Lingampalli, S. R., Manjunath, K., Shenoy, S., Waghmare, U. V., & Rao, C. N. R. (2016).
Zn2NF and related analogues of ZnO. Journal of the American Chemical Society, 138,
8228–8234.
85. Prakash, T. (2012). Review on nanostructured semiconductors for dye sensitized solar cells.
Electronic Materials Letters, 8, 231–243.
86. Bale, C. W., Bélisle, E., Chartrand, P., Decterov, S. A., Eriksson, G., Gheribi, A. E., Hack, K.,
Jung, I. H., Kang, Y. B., Melançon, J., Pelton, A. D., Petersen, S., Robelin, C., Sangster, J.,
Spencer, P., & Van Ende, M. A. (2016). Reprint of: FactSage thermochemical software and
databases, 2010–2016. Calphad, 55, 1–19.
87. Li, Y., George, J., Liu, X., & Dronskowski, R. (2015). Synthesis, structure determination and
electronic structure of magnesium nitride chloride, Mg2NCl. Zeitschrift für Anorganische und
Allgemeine Chemie, 641, 266–269.
88. Brogan, M. A., Hughes, R. W., Smith, R. I., & Gregory, D. H. (2012). Structural studies of
magnesium nitride fluorides by powder neutron diffraction. Journal of Solid State Chemistry,
185, 213–218.
89. Stoumpos, C. C., Malliakas, C. D., & Kanatzidis, M. G. (2013). Semiconducting tin and lead
iodide perovskites with organic cations: Phase transitions, high mobilities, and near-infrared
photoluminescent properties. Inorganic Chemistry, 52, 9019–9038.
90. Lyberis, A., Patriarche, G., Gredin, P., Vivien, D., & Mortier, M. (2011). Origin of light scat-
tering in ytterbium doped calcium fluoride transparent ceramic for high power lasers. Journal
of the European Ceramic Society, 31, 1619–1630.
91. Al-Azzawi, M. A. (2016). Synthesis and characterization of single crystalline metal nitride
fluorides. Youngstown State University.
92. Vikse, K. L., Woods, M. P., & McIndoe, J. S. (2010). Pressurized sample infusion for the con-
tinuous analysis of air- and moisture-sensitive reactions using electrospray ionization mass
spectrometry. Organometallics, 29, 6615–6618.
93. Smolentsev, S., Li, F. C., Morley, N., Ueki, Y., Abdou, M., & Sketchley, T. (2013). Construction
and initial operation of MHD PbLi facility at UCLA. Fusion Engineering and Design, 88,
317–326.
94. Sanz, M. (2010). A system for cooling inside a glove box. Journal of Chemical Education,
87, 854–855.
95. Werner, K., Pommer, L., & Broström, M. (2014). Thermal decomposition of hemicelluloses.
The Journal of Analytical and Applied Pyrolysis, 110, 130–137.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 517

96. Roskosz, M., Bouhifd, M. A., Jephcoat, A. P., Marty, B., & Mysen, B. O. (2013). Nitrogen
solubility in molten metal and silicate at high pressure and temperature. Geochimica et
Cosmochimica Acta, 121, 15–28.
97. Maliyekkal, S.  M., Anshup, Antony, K.  R., & Pradeep, T. (2010). High yield combustion
synthesis of nanomagnesia and its application for fluoride removal. Science of The Total
Environment, 408, 2273–2282.
98. Dmitrienko, S. G., Medvedeva, O. M., Ivanov, A. A., Shpigun, O. A., & Zolotov, Y. A. (2002).
Determination of gallic acid with 4-nitrobenzenediazonium tetrafluoroborate by diffuse
reflectance spectrometry on polyurethane foam. Analytica Chimica Acta, 469, 295–301.
99. Yang, S., Feng, X., Zhi, L., Cao, Q., Maier, J., & Müllen, K. (2010). Nanographene-­
constructed hollow carbon spheres and their favorable electroactivity with respect to lithium
storage. Advanced Materials, 22, 838–842.
100. Dong, S., Chen, X., Gu, L., Zhou, X., Xu, H., Wang, H., Liu, Z., Han, P., Yao, J., Wang, L.,
Cui, G., & Chen, L. (2011). Facile preparation of mesoporous titanium nitride microspheres
for electrochemical energy storage. ACS Applied Materials & Interfaces, 3, 93–98.
101. Maeda, K., Kuriki, R., Sekizawa, K., & Ishitani, O. (Eds.). (2015). Metal-complex/semi-
conductor hybrids for carbon dioxide fixation. In SPIE optics + photonics for sustainable
energy. SPIE.
102. Smieja, J. M., & Kubiak, C. P. (2010). Re(bipy-tBu)(CO)3Cl−improved catalytic activity for
reduction of carbon dioxide: IR-spectroelectrochemical and mechanistic studies. Inorganic
Chemistry, 49, 9283–9289.
103. Sakata, Y., Matsuda, Y., Yanagida, T., Hirata, K., Imamura, H., & Teramura, K. (2008). Effect
of metal ion addition in a Ni supported Ga2O3 photocatalyst on the photocatalytic overall
splitting of H2O. Catalysis Letters, 125, 22–26.
104. Kubota, J., & Domen, K. (2013). Photocatalytic water splitting using oxynitride and nitride
semiconductor powders for production of solar hydrogen. The Electrochemical Society
Interface, 22, 57–62.
105. Maeda, K., Teramura, K., & Domen, K. (2008). Effect of post-calcination on photocatalytic
activity of (Ga1−xZnx)(N1−xOx) solid solution for overall water splitting under visible light.
Journal of Catalysis, 254, 198–204.
106. Habazaki, H., Konno, Y., Aoki, Y., Skeldon, P., & Thompson, G. E. (2010). Galvanostatic
growth of nanoporous anodic films on iron in ammonium fluoride−ethylene glycol electro-
lytes with different water contents. Journal of Physical Chemistry C, 114, 18853–18859.
107. Kasahara, A., Nukumizu, K., Takata, T., Kondo, J. N., Hara, M., Kobayashi, H., & Domen,
K. (2003). LaTiO2N as a visible-light (≤600 nm)-driven photocatalyst (2). The Journal of
Physical Chemistry B, 107, 791–797.
108. Le Paven-Thivet, C., Ishikawa, A., Ziani, A., Le Gendre, L., Yoshida, M., Kubota, J., Tessier,
F., & Domen, K. (2009). Photoelectrochemical properties of crystalline perovskite lantha-
num titanium oxynitride films under visible light. Journal of Physical Chemistry C, 113,
6156–6162.
109. Vidysagar, D., Ghugal, S.  G., Kulkarni, A., Mishra, P., Shende, A.  G., Jagannath, Umare,
S., & Sasikala, R. (2018). Silver/Silver(II) oxide (Ag/AgO) loaded graphitic carbon nitride
microspheres: An effective visible light active photocatalyst for degradation of acidic dyes
and bacterial inactivation. Applied Catalysis B: Environmental, 221, 339–348.
110. Schipper, D. E., Zhao, Z., Leitner, A. P., Xie, L., Qin, F., Alam, M. K., Chen, S., Wang, D.,
Ren, Z., Wang, Z., Bao, J., & Whitmire, K. H. (2017). ACS Nano, 11, 4051–4059.
111. Hou, Y., Qiu, M., Zhang, T., Zhuang, X., Kim, C. S., Yuan, C., & Feng, X. (2017). Ternary
porous cobalt phosphoselenide nanosheets: An efficient electrocatalyst for electrocatalytic
and photoelectrochemical water splitting. Advanced Materials, 29, 1701589.
112. Yu, X., Zhao, Z., Sun, D., Ren, N., Ding, L., Yang, R., Ji, Y., Li, L., & Liu, H. (2018). TiO2/
TiN core/shell nanobelts for efficient solar hydrogen generation. Chemical Communications,
54, 6056–6059.
518 S. Kumar et al.

113. Ohashi, T., Wang, Y., & Shimada, S. (2010). Preparation and high catalytic performance
of hollow BN spheres-supported Ni for hydrogen production from methanol. Journal of
Materials Chemistry, 20, 5129–5135.
114. He, Z., Kim, C., Jeon, T.  H., & Choi, W. (2018). Hydrogenated heterojunction of boron
nitride and titania enables the photocatalytic generation of H2 in the absence of noble metal
catalysts. Applied Catalysis B: Environmental, 237, 772–782.
115. Yang, X. J., Li, L. L., Sang, W. L., Zhao, J. L., Wang, X. X., Yu, C., Zhang, X. H., & Tang,
C. C. (2017). Boron nitride supported Ni nanoparticles as catalysts for hydrogen generation
from hydrolysis of ammonia borane. Journal of Alloys and Compounds, 693, 642–649.
116. Qiu, X., Wu, X., Wu, Y., Liu, Q., & Huang, C. (2016). The release of hydrogen from ammonia
borane over copper/hexagonal boron nitride composites. RSC Advances, 6, 106211.
117. Arunachalam, M., Yun, G., Ahn, K.-S., Seo, W.-S., Jung, D. S., & Kang, S. H. (2018). Unique
photoelectrochemical behavior of TiO2 nanorods wrapped with novel titanium Oxy-Nitride
(TiOxNy) nanoparticles. International Journal of Hydrogen Energy, 43, 16458–16467.
118. Ravi, L., Boopathi, K., Panigrahi, P., & Krishnan, B. (2018). Growth of gallium nitride
nanowires on sapphire and silicon by chemical vapor deposition for water splitting applica-
tions. Applied Surface Science, 449, 213–220.
119. Dholam, R., Patel, N., & Miotello, A. (2011). Efficient H2 production by water-splitting using
indium tin-oxide/V-doped TiO2 multilayer thin film photocatalyst. International Journal of
Hydrogen Energy, 36, 6519–6528.
120. Huang, C., Chen, C., Zhang, M., Lin, L., Ye, X., Lin, S., Antonietti, M., & Wang, X. (2015).
Carbon-doped BN nanosheets for metal-free photoredox catalysis. Nature Communications,
6, 7698.
121. Wu, Y., Wu, X., Liu, Q., Huang, C., & Qiu, X. (2017). Magnetically recyclable Ni@h-BN
composites for efficient hydrolysis of ammonia borane. International Journal of Hydrogen
Energy, 42, 16003–16001.
122. Zhang, H., Tong, C.-J., Zhang, Y., Zhanga, Y.-N., & Liu, L.-M. (2015). Porous BN for hydro-
gen generation and storage. Journal of Materials Chemistry A, 3, 9632.
123. Lv, Z., Zhang, Z., Lu, L., Liu, J., & Song, W. (2018). Manipulation structure of carbon nitride
via trace level Iron with improved interfacial redox activity and charge separation for syn-
thetic enhancing photocatalytic hydrogen evolution. Applied Surface Science, 456, 609–614.
124. Wen, J., Xie, J., Chen, X., & Li, X. (2017). A review on g-C3N4-based photocatalysts. Applied
Surface Science, 391, 72–123.
125. Luo, B., Song, R., Geng, J., Jing, D., & Huang, Z. (2019). Strengthened spatial charge sepa-
ration over Z-scheme heterojunction photocatalyst for efficient photocatalytic H2 evolution.
Applied Surface Science, 475, 453–461.
126. Fu, J., Xu, Q., Low, J., Jiang, C., & Yu, J. (2019). Ultrathin 2D/2D WO3/g-C3N4 step-scheme
H2-production photocatalyst. Applied Catalysis B: Environmental, 243, 556–565.
127. Liu, X., Xu, S., Chi, H., Xu, T., Guo, Y., Yuan, Y., & Yang, B. (2019). Ultrafine 1D graphene
interlayer in g-C3N4/graphene/recycled carbon fiber heterostructure for enhanced photocata-
lytic hydrogen generation. Chemical Engineering Journal, 359, 1352–1359.
128. Wang, J., Liu, D., Liu, Q., Peng, T., Li, R., & Zhou, S. (2019). Effects of the central metal
ions on the photosensitization of metalloporphyrins over carbon nitride for visible-light-­
responsive H2 production. Applied Surface Science, 464, 255–261.
129. Liu, J., Fang, W., Wei, Z., Qin, Z., Jiang, Z., & Shangguan, W. (2018). Efficient photocata-
lytic hydrogen evolution on N-deficient g-C3N4 achieved by a molten salt post-treatment
approach. Applied Catalysis B: Environmental, 238, 465–470.
130. Zhao, D., Wang, M., Kong, T., Shang, Y., Du, X., Guo, L., & Shen, S. (2019). Electronic
pump boosting photocatalytic hydrogen evolution over graphitic carbon nitride. Materials
Today Chemistry, 11, 296–302.
131. Mo, Z., Xu, H., She, X., Song, Y., Yan, P., Yi, J., Zhu, X., Lei, Y., Yuan, S., & Li, H. (2019).
Constructing Pd/2D-C3N4 composites for efficient photocatalytic H2 evolution through
nonplasmon-­induced bound electrons. Applied Surface Science, 467–468, 151–157.
19  Metal Nitrides and Graphitic Carbon Nitrides as Novel Photocatalysts… 519

132. Boppella, R., Yang, W., Tan, J., Kwon, H.-C., Park, J., & Moon, J. (2019). Black phosphorus
supported Ni2P co-catalyst on graphitic carbon nitride enabling simultaneous boosting charge
separation and surface reaction. Applied Catalysis B: Environmental, 242, 422–430.
133. Sun, S., Li, J., Cui, J., Gou, X., Yang, Q., Jiang, Y., Liang, S., & Yang, Z. (2019).
Simultaneously engineering K-doping and exfoliation into graphitic carbon nitride (g-C3N4)
for enhanced photocatalytic hydrogen production. International Journal of Hydrogen Energy,
44, 778–787.
134. Bai, J., Yin, C., Xu, H., Chen, G., Ni, Z., Wang, Z., Li, Y., Kang, S., Zheng, Z., & Li,
X. (2018). Facile urea-assisted precursor pre-treatment to fabricate porous g-C3N4 nanosheets
for remarkably enhanced visible-light-driven hydrogen evolution. Journal of Colloid and
Interface Science, 532, 280–286.
135. Wang, H., Bian, Y., Hu, J., & Dai, L. (2018). Highly crystalline sulfur-doped carbon nitride
as photocatalyst for efficient visible-light hydrogen generation. Applied Catalysis B:
Environmental, 238, 592–598.
136. Yuan, Y.-J., Shen, Z., Wu, S., Su, Y., Pei, L., Ji, Z., Ding, M., Bai, W., Chen, Y., Yu, Z.-T.,
& Zou, Z. (2019). Liquid exfoliation of g-C3N4 nanosheets to construct 2D-2D MoS2/g-­
C3N4 photocatalyst for enhanced photocatalytic H2 production activity. Applied Catalysis B:
Environmental, 246, 120–128.
137. Wang, J., Yang, Z., Yao, W., Gao, X., & Tao, D. (2018). Defects modified in the exfoliation
of g-C3N4 nanosheets via a self-assembly process for improved hydrogen evolution perfor-
mance. Applied Catalysis B: Environmental, 238, 629–637.
138. Li, H., Deng, P., & Hou, Y. (2018). Cobalt disulfide/graphitic carbon nitride as an efficient
photocatalyst for hydrogen evolution reaction under visible light irradiation. Materials
Letters, 229, 217–220.
139. Peng, W., Zhang, S.-S., Shao, Y.-B., & Huang, J.-H. (2018). Bimetallic PtNi/g-C3N4 nano-
tubes with enhanced photocatalytic activity for H2 evolution under visible light irradiation.
International Journal of Hydrogen Energy, 43, 22215–22225.
140. Liu, Y., Xu, X., Zhang, J., Zhang, H., Tian, W., Li, X., Tade, M.  O., Sun, H., & Wang,
S. (2018). Flower-like MoS2 on graphitic carbon nitride for enhanced photocatalytic and
electrochemical hydrogen evolutions. Applied Catalysis B: Environmental, 239, 334–344.
141. Bhunia, K., Chandra, M., Khilari, S., & Pradhan, D. (2019). Bimetallic PtAu alloy
nanoparticles-­integrated g-C3N4 hybrid as an efficient photocatalyst for water-to-hydrogen
conversion. ACS Applied Materials & Interfaces, 11, 478–488.
142. Wang, X., Zhao, Z., Shu, Z., Chen, Y., Zhou, J., Li, T., Wang, W., Tan, Y., & Sun, N. (2018).
One-pot synthesis of metakaolin/g-C3N4 composite for improved visible-light photocatalytic
H2 evolution. Applied Clay Science, 166, 80–87.
Chapter 20
Highly Functionalized Nanostructured
Titanium Oxide-Based Photocatalysts
for Direct Photocatalytic Decomposition
of NOx/VOCs

Katchala Nanaji, Manavalan Vijayakumar, Ammaiyappan Bharathi Sankar,


and Mani Karthik

20.1  Introduction

In a new civilized world, air pollution is one of the major concerns, and it has a seri-
ous toxic impact on human health as well as the environment. The ground-level
ozone (GLO), particulate matter (PM), sulphur oxides (SOx), carbon monoxide
(CO), nitrogen oxides (NOx) and volatile organic compounds (VOCs) are consid-
ered as six major air pollutants according to the World Health Organization (WHO).
Air pollutants can create several serious diseases like asthma, eye irritation, head-
aches, skin diseases, respiratory and cardiovascular diseases, foetal growth, ven-
tricular hypertrophy, lung cancer, psychological complications and low birth weight
[1–4]. Combination of hydrocarbons and NOx in the atmosphere produces a photo-
chemical smog via photochemical oxidation under the sunlight, and the photochem-
ical smog is environmentally hazardous. Generally, air pollutants can be created by
natural sources or manmade sources, and it is classified as primary air pollutants or
secondary air pollutants. Primary air pollutants are directly emitted from natural
volcanic eruption, vehicle exhaust and industrial processes. Secondary air pollut-
ants are not directly emitted from the direct resources, but primary pollutants could

K. Nanaji · M. Karthik (*)


Centre for Nanomaterials, International Advanced Research Centre for Powder Metallurgy
and New Materials (ARCI), Balapur, Hyderabad, Telangana, India
e-mail: mkarthik@project.arci.res.in
M. Vijayakumar
Centre for Nanomaterials, International Advanced Research Centre for Powder Metallurgy
and New Materials (ARCI), Balapur, Hyderabad, India
Global Innovative Centre for Advanced Nanomaterials (GICAN), Collage of Science,
Engineering and Environment, The University of Newcastle, NSW, Callaghan, Australia
A. B. Sankar
School of Electronics Engineering, Vellore Institute of Technology (VIT), Chennai Campus,
Chennai, Tamil Nadu, India

© Springer Nature Switzerland AG 2021 521


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_20
522 K. Nanaji et al.

react or interact to form secondary pollutants. For example, ground-level ozone is


one of the secondary air pollutants which formed from the photochemical smog.

20.1.1  Sources of Air Pollutants and Its Impact

PM pollutants can be created by manmade sources and/or natural sources. For


example, forest and grassland fires, natural volcanoes, living vegetation, sea spray,
etc. are the major resources of PM pollutants. On the other hand, PM pollutants can
also be created by manmade activities such as burning of fossil fuels in the vehicles,
power plants and various industrial activities. The creation of fine particles in the air
can cause several health hazards to the living things. On the other hand, the emis-
sions of NOx and VOCs are also creating severe impacts on human health, environ-
ment and biological ecosystem [4–7]. The emissions of these pollutants cause
severe environmental threats including acid rain, PM  releases, smog and green-
house gasses. VOCs can react with NOx and other airborne chemicals in the air to
produce the ozone through photochemical oxidation under the sunlight, and this
ozone is a primary component of smog as environmentally hazardous. Figure 20.1
illustrates the various potential resources of major air pollutants.
Due to air pollution, around 8.3 million people are killed in the world, and about
87% of death rate occurs in the low- and middle-income countries as per the WHO
reports (WHO, 2016) [1, 2]. For example, India and China are among the most pol-
luted countries owing to the rapid urbanization [1, 2]. The estimated number of
premature pollution-related deaths per year in the world is represented in Fig. 20.2.
Photocatalytic oxidation (PCO) using titanium dioxide (TiO2) semiconductors
has been recognized as one of the efficient and economic techniques for the com-
plete decomposition of harmful pollutants into nontoxic final products [8, 9]. The
novel photocatalytic processes are being developed that are totally a safer and
cleaner process and efficient and environmental-friendly nature. Generally, the pho-
tocatalytic processes can be utilized not only in ultraviolet but also in solar or visible
light region. The functionalized photocatalytic systems will be essential for the
future prosperity of mankind.
In this context, several studies have been conducted by using TiO2 photocatalyst
for the photocatalytic oxidation of various pollutants [10–18]. However, most of the
reports and reviews have mainly described on single pollutant treatment system
using either TiO2 or TiO2-supported catalysts. For example, Anpo and co-workers
have extensively studied and also reviewed [13–15] the photocatalytic degradation
of NOx by using TiO2 as well as modified TiO2 as photocatalysts. However, there
are very limited studies on the photocatalytic removals of multiple air pollutants
over TiO2-based photocatalysts. Generally, indoor environment in most cases will
often contain more than one pollutant. Hence, the photocatalytic material capable of
destroying multi-air pollutants in the single system is showed more attraction for
practical photocatalytic implementation. On the other hand, the dual functional
adsorbent/photocatalyst system has been suggested as one of the effective and
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 523

Fig. 20.1  Various potential resources of air pollutants such as particulate matter (PM), nitrogen
oxides (NOx) and volatile organic compounds (VOCs)

economic techniques for the removal of pollutants [19–21]. From the obtained
results, it was found that the removal efficiency of pollutant over combination of
adsorption/desorption with photocatalytic oxidation system was three times higher
as compared with the normal photocatalytic oxidation process. Thus, these adsor-
bent/photocatalyst composite materials deserve research attention due to their com-
bination properties of adsorption and photocatalysts which enhance the activity and
selectivity towards specific air pollutants.
In the present chapter, the photocatalytic destructions of major air pollutants like
VOCs and NOx over TiO2 and functionalized nanostructured TiO2 photocatalysts
are discussed. The fundamental photocatalytic concepts and strategies of photocata-
lysts on environmental remediation are summarized. This chapter also provides
some highlights and examples regarding recent progressive research on dual func-
tional composites of TiO2/porous catalytic adsorbents for photocatalytic destruction
of air pollutants. Finally, the current and future research directions on photocatalytic
decomposition of NOx and VOCs for environmental remediation are also
highlighted.
524 K. Nanaji et al.

Fig. 20.2  The estimated number of premature pollution-related deaths per year in the world [1, 2]

20.2  P
 hotocatalytic Destruction of Air Pollutants by Using
TiO2 Photocatalysts

Photocatalytic destruction of air pollutants using semiconductor photocatalyst is


one of the most promising techniques owing to its nontoxicity, long-term durability,
high photosensitivity, environmental-friendly nature, easy availability, economics
and high reactivity for the complete decomposition of toxic compounds into harm-
less products. In general, semiconductors like TiO2, ZnO, WO3, CdS, Fe2O3, ZnS
and SnO2 are commonly utilized as photocatalysts. Among various semiconductor
photocatalysts investigated, TiO2 is one of the most effective and widely used com-
mercial photocatalysts. And TiO2 photocatalyst has an outstanding advantage and
strong capacity in the mineralization of environmental pollutants even with a low
concentration. Thus the photocatalytic oxidation of air contaminants using TiO2
photocatalysts has been extensively studied by several research groups over the past
several decades [4, 16–18]. Although TiO2 photocatalysts have been extensively
investigated for environmental remediation, the large band gap of anatase TiO2
(3.2 eV) limits its visible light absorption (λ < 400 nm). Thus, the different tech-
niques have been identified to improve the visible light absorption of TiO2 photo-
catalyst via modification of the photocatalyst by doping or nanoparticle loading,
non-metal doping, heterojunction with other semiconductors sensitization and
introduction of defect sites. The various percentage distributions of already reported
data on modified TiO2-based photocatalysts for decomposition of air pollutants are
depicted in Fig. 20.3.
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 525

Fig. 20.3  The various percentage distributions of already reported data on modified TiO2-based
photocatalysts for decomposition of air pollutants [4]

20.3  P
 hotocatalytic Destruction of Air Pollutants by TiO2
and TiO2-Supported Photocatalysts

20.3.1  TiO2 Semiconductor as Photocatalysts

TiO2 is one of the potential candidates for destruction of environmental pollutants to


diminish the serious and detrimental effects of air and water pollutions. TiO2 photo-
catalyst has an outstanding advantage and strong capacity in the mineralization of
environmental pollutants with a low concentration. Hence, TiO2 photocatalysts have
been widely studied and applied in many potential applications owing to their high
efficiency, economics, non-toxicity, photostability and sensitivity. Numerous stud-
ies on the photodecomposition of pollutants by using these photocatalysts have been
conducted in the last few decades [8–12], and the preparation of highly active pho-
tocatalysts have been extensively reported [13–15]. From the several studies, it was
found that the photocatalytic efficiency of TiO2 depends on specific surface area,
crystalline size and morphology of the photocatalysts.
Anatase, rutile and brookite are three crystalline phases of TiO2 photocatalysts.
The metastable phase anatase is transformed easily into the rutile phase upon higher
temperature (600–800  °C) treatment. TiO2 with anatase phase has high chemical
stability and is readily available and highly photoactive for oxidation processes. The
526 K. Nanaji et al.

anatase form of TiO2 has 3.2 eV of energy band gap with wavelength that corre-
sponds to UV light about 388 nm. The 3.2 eV band gap is suitable for a wide range
of readily available lamps, but it is not an ideal candidate for solar applications. The
band gap of rutile is around 3.0 eV with wavelength that corresponds to violet light
about 413 nm, but few studies only reported on photocatalytic performance of rutile
photocatalyst [22]. Generally, it is observed that the rutile form has a less photocata-
lytic activity than that of anatase. Most of the reports have shown that rutile is pos-
sessing a poor photocatalytic reactivity [22]. However, a small mixture of rutile with
anatase showed higher photocatalytic performance as compared with the pure ana-
tase form [22, 23]. The optimum compositions of rutile with anatase mixture as a
photocatalyst have not been yet studied. It has been noted that TiO2 photocatalyst
has a less removal efficiency which is not sufficient for practical implementations.
In addition, TiO2 photocatalyst generally has a low adsorption capacity of pollutant
molecules. Therefore, the design and development of multifunctional catalysts (vis-
ible light responsive, high reactivity and high adsorptivity) will be important for the
practical applications. Based on the above considerations, the development of sup-
ported TiO2 photocatalysts by using various supports such as porous glass, silica,
zeolites and zeo-type molecular sieves have been extensively reported [19–21].
The highly dispersed and well-defined TiO2 particles supported or anchored onto
the various inert supports are increasing research concerns due to their significant
and high photocatalytic performance and selectivity as compared with bulk TiO2
photocatalyst. Thus, design and development of highly efficient photocatalysts
incorporated/encapsulated zeolite-based molecular sieves have attracted research
interests in many research areas [24–26]. The photocatalytic activities of well-­
dispersed TiO2 particles supported/anchored onto the zeolites and zeo-type molecu-
lar sieves can enhance the photocatalytic activity and selectivity owing to the
substantial changes in their electronic properties [19–21]. The schematic represen-
tation of simultaneous adsorption and photocatalytic decomposition of TiO2-­
supported photocatalyst is depicted in Fig. 20.4.
In the following section, the photocatalytic destruction of air pollutants like NOx
and VOCs using TiO2-supported zeolite photocatalysts under UV irradiation will be
elaborated. Furthermore, the photocatalytic activity of titanium/titanium oxide-­
functionalized mesoporous-based photocatalysts for destruction of air pollutants
will also be summarized.

20.3.2  H
 ighly Dispersed Titanium Dioxides onto
Various Supports

It was reported that highly dispersed TiO2 onto porous Vycor glass has showed
higher photocatalytic performance as compared with bulk TiO2 photocatalysts [25,
26]. Anpo et al. [27] have highlighted that TiO2 anchored onto silica glass was syn-
thesized by metal ion implantation technique using TiCl4 as titanium precursor. The
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 527

Fig. 20.4  Schematic illustration of simultaneous adsorption and photocatalytic decomposition of


air pollutants by using TiO2-supported photocatalyst

prepared catalysts were successfully utilized for the photocatalytic degradations of


an aqueous 2-propanol solution in the liquid-phase reaction, and it was found that
TiO2 anchored onto silica glass showed higher photocatalytic performance as com-
pared with bulk TiO2 catalysts. From the above studies, it can be easily observed
that highly dispersed TiO2 onto various supports has showed higher photocatalytic
performance as compared with bulk TiO2 photocatalysts. In order to understand the
role of supported materials on the photocatalytic activity, the structure and composi-
tions of supported materials, particularly zeolites and zeo-type materials as sup-
ported materials, are discussed in the following section.
528 K. Nanaji et al.

20.3.3  M
 icroporous Zeolites as Supports
for Photocatalytic Destruction
Structure and Composition of the Microporous Zeolites

Zeolites and zeo-type molecular sieves are the most widely applicable materials not
only in catalysis but also in adsorption, separation and environmental pollution con-
trol due to their special features such as easy tailoring of acidity and porosity, shape
selectivity, hydrophilic and hydrophobic properties, high hydrothermal and thermal
stability and eco-friendly nature [28]. Zeolites are three-dimensional crystalline alu-
minosilicate materials. The general structural formula of the zeolites is represented
as Mx/nO.xAl2O3.ySiO2, where n is the valance of the cation (M), x and y are the
total number of tetrahedral per unit cell and y/x is the atomic Si/Al ratio which vary
from a minimal value of 1 (as per Lowenstein rule) to infinite. 3D arrangement of
AlO4− and SiO4 tetrahedral is connecting via oxygen linkage to form subunits and
repeating identical building blocks (unit cells). Each and every aluminium atom
incorporated into the framework of zeolite can create a negative charge and that
negative charge can be balanced by exchangeable cations in the cages and/or chan-
nels of zeolites. Zeolites are interconnected with bridged hydroxyl groups (Si(OH)
Al) via oxygen linkage with tetrahedral Si and Al atoms. The generation of Bronsted
and Lewis acid sites in zeolites upon calcination is depicted in Fig. 20.5.
Bronsted acid sites of zeolites can be formed by bridged hydroxyl groups, and
Lewis acidity can be created in zeolites by incorporating Lewis acidic metal cations
as charge-compensating species instead of protons. Further strong Lewis acid sites
are created when there are defects in the aluminosilicate framework, and unsatu-
rated trigonal aluminium cations are generated during calcination of zeolites in
higher temperature [29, 30]. The attractive properties of zeolites or zeo-type materi-
als are most essential for the design of well-dispersed transition metal oxides such
as Ti, Mo, Cu, Co, V, Mn, Cr, etc., within the zeolite framework/cavities, which
excited under UV irradiation through the electron charge transfer as shown in
Fig. 20.6.

Photocatalytic Decomposition of NOx Using Zeolite-Based Photocatalysts

Among various air contaminants, NOx is one of the most common air pollutants
which are found in most of the indoor and/or outdoor environment. There is a spe-
cial interest in designing the titanium/titanium oxide-supported zeolites because
these fascinating composite materials are possessing significantly higher photocata-
lytic reactivity and selectivity as compared to bulk TiO2 catalysts. It was found that
highly dispersed TiO2 particles onto the cavities of zeolite exhibited much higher
photocatalytic decomposition of NOx into N2 and O2 as compared with bulk TiO2
photocatalyst [29, 30]. It was reported on the photoluminescence properties of the
titanium silicalite (TS-2) catalyst for the direct decomposition of NO into N2, N2O
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 529

Fig. 20.5  Schematic representation of Bronsted and Lewis acid sites in zeolites

Mesoporous silica

e- e- e- e-
Conducon Band
e- e-
O2- hν O2-
(250- 300 nm)
Mn+ M(n-1)+
Excitaon

O2- O2- O2- O2- O2- O2-


hν Eg
Light irradiaon (~ 380 nm)
Zeolite
Valence Band Single-site Charge transfer
h+ h+ h+ h + h+ h +
photocatalyst excited state

Semiconducng bulk TiO2


O2- O2- O2-

Ti4+ V5+ Cr6+


O2- O2- O2- O2- O2- O2- O2- O2- O2-

Fig. 20.6  The electron charge transfers excited states of transition metal oxide species under UV
irradiation [29, 30]
530 K. Nanaji et al.

and O2 under UV irradiation at 295 K [31]. Anpo et al. [32–34] have found that the
tetrahedrally coordinated Ti with high dispersion exhibited much higher photocata-
lytic activity for NO decomposition as compared to bulk TiO2 catalysts. It was
observed from the literature that N2O and NO2 are formed as major products, while
N2 and O2 are scarcely formed under the absence of O2 [35]. Courbon and Pichat
[36] have found entirely different observation, and they reported that the photocata-
lytic decomposition of NOx over powdered TiO2 showed the predominant product
of N2O rather than the formation of N2 and O2.
Anpo and co-workers have reported that TiO2 anchored on Y zeolite via ion-­
exchange technique showed higher photocatalytic decomposition of NOx into N2,
O2 and N2O with a high selectivity of N2 [35]. The high dispersion of TiO2 in the
cavities of zeolite has exhibited a remarkable selectivity of N2 in the decomposition
of NOx as well as superior activity than TiO2 or TiO2 photocatalyst on the surface of
silica/zeolites obtained by a simple impregnation technique [37, 38]. Anpo and co-­
workers have also highlighted that the tetrahedrally coordinated highly dispersed
titanium dioxides were prepared on the surface of microporous Vycor glass and the
prepared catalyst showed higher photocatalytic performance than that of the bulk
TiO2 catalyst [25, 26, 38].
Hu et  al. [39] proposed that TiO2-incorporated zeolites (TiO2/zeolite) were pre-
pared by an ion-exchange method and their photocatalytic activities were examined.
Then the photocatalytic performance of TiO2 prepared by different methods was com-
pared with bulk TiO2 for the photocatalytic decomposition of NO at 295 K as depicted
in Fig. 20.7. It can be clearly seen from the figure that the selectivity of N2 strongly
depends on the various types of the Ti/zeolite-based photocatalysts. The yield of N2 in
the decomposition reaction was found to be much higher in low Si/Al ratios as com-
pared with zeolites with high Si/Al ratios. The exchanged Ti/zeolites showed higher
selectivity of N2 than impregnated Ti/zeolites. The selectivity of N2O was found to be

Fig. 20.7  Comparison of the yield of N2 and N2O from the direct photocatalytic decomposition of
NO over Ti-oxide/zeolite-based catalysts prepared in different methods [39]. Numbers in parenthe-
ses indicate Si/Al ratio of zeolites: aTi content 10 wt.% and bTi content 1.1 wt.%
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 531

the major product on the bulk TiO2 photocatalyst as well as on the impregnated Ti/
zeolite photocatalysts. Hence, the obtained results clearly demonstrated that the pho-
tocatalytic activity and selectivity of N2 were decreased when the aggregated titanium
oxides in the impregnated Ti/zeolite photocatalysts were increased.

Reaction Pathways for Photocatalytic Abatement of NOx

The photocatalytic abatement of NOx might undergo many reaction steps. The pho-
toelectrons (e−) and positively charged holes (h+) are generated on photocatalyst
surface upon light irradiation. For the typical photocatalytic reaction, the generation
of electron-hole (e−, h+) pairs, hydroxyl radicals (•OH) and oxygen radicals (O2•−) on
the photocatalyst surface can play an important role. The following reaction steps
are carried out on the photocatalytic decomposition of NOx under the sunlight [22]:
hv
TiO2 → ( TiO2 ) → h + + e − (20.1)

h + + H 2 O → OH∗ + H + (20.2)

H + + e − + O2 → HO2∗ (20.3)

2HO2− → H 2 O2 + O2 (20.4)

2e − + O2 + 2H + → H 2 O2 (20.5)

2 h + + 2H 2 O → H 2 O2 + 2H + (20.6)

NO + HO2 → NO2 + OH − (20.7)

NO2 + NO + H 2 O → 2HONO (20.8)

NO2 + OH − → HNO3 (20.9)

NO + OH − → HNO2 → H + + NO2− (20.10)

HNO2 + OH − → NO2 + H 2 O (20.11)

At the air/semiconductor interface, the basic reaction is taking place in the presence
of H2O, O2, NO and NO2 which are leading to the nitrate and nitrite formation. The
possible reaction pathways of the photocatalytic decomposition of NO on the sur-
face of TiO2 photocatalyst are illustrated in Fig. 20.8. Under light irradiation, the
electron transfers happened from the electron-trapped centres into the antibonding
orbitals of adsorbed NO molecules, and then the formation of N(ads) and O(ads)
surface species occurred. Then, the above species would migrate on the surface of
TiO2, and then it can react with other surface species such as NO(ads), N(ads) and
O(ads) to form several by-products such as N2O(gas), NO2(ads), O2(gas) and
N2(gas). It is observed that the primary reaction is NO(ads) + N(ads) → N2O(ads),
which generates N2O as the major product. On the other hand, the mechanism of
532 K. Nanaji et al.

Fig. 20.8  The possible reaction pathways of photocatalytic destruction of NO on the surface of
TiO2 photocatalyst [40–42]

photodecomposition of NO(ads) to N(ads) + O(ads) via NO (ads) (intermediate) is


also proposed [40–42].

Photocatalytic Oxidation of VOCs Using Zeolite-Based Photocatalysts

The photodegradation mechanism of VOCs and NOx has been extensively investi-
gated. In the photodecomposition process, TiO2 can behave as an electron donors or
acceptors when the molecules can reach the surface of the photocatalyst. The posi-
tive holes and electrons are very active species which participate in redox reactions
on the photocatalyst surface with water, oxygen, adsorbed organic or inorganic
compounds which leads to the degradation of the pollutants. The electrons (e−) and
holes (h+) can directly react with adsorbed organic or inorganic compounds on the
surface of the photocatalyst. The reaction of h+ and e− with water molecules and
oxygen leads to the intermediate formation which generates a very powerful radi-
cals like hydroxyl (OH•) and superoxide (O2−•) radicals that are responsible for the
redox processes at the semiconductor surface in order to promote the pollutant deg-
radation as represented in Fig. 20.9.
However, the recombination of positively charged holes (h+) and electrons (e-)
rapidly occurred, and hence there is no reaction that takes place. Therefore, major
efforts are needed to inhibit recombination of charge carriers in order to promote the
photocatalytic activity. In this regard, the mixed phases of anatase and rutile TiO2
can strongly promote the charge separation via trapping of holes and electrons
across the various crystal phases. The electron acceptors and photo-generated elec-
trons should react to avoid recombination. The charge separation is also achieved by
doping of TiO2 with metals and non-metals. Furthermore, doped photocatalyst can
use the entire solar spectral radiation, and it acts as a visible light photocatalyst.
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 533

Fig. 20.9  Photocatalytic processes on semiconductor photocatalysts [22]

Fig. 20.10  The basic principle of photocatalytic oxidation of VOCs over metal- and non-metal-­
doped TiO2 [43]

Generally, the solar spectrum consists of only 5–8% of UV, but it consists of 43%
visible light which enables the activation of the semiconductors, e.g. TiO2 at longer
wavelengths (red shift, λ > 380 nm). The basic principle of photocatalytic destruc-
tion of VOCs is depicted in Fig. 20.10.
The reaction steps involved in the photocatalytic destruction of VOC pollutants
over semiconductor TiO2 photocatalyst are as follows [43]:

Photoexcitation : TiO2 + hϑ → h + + e − (20.12)

Oxidation reaction : OH − + h + → OH• (20.13)

Reduction reaction : O2(ads) + e − → O2−(ads)


(20.14)
534 K. Nanaji et al.

Ionization of water : H 2 O → OH − + H + (20.15)

Protonation of superoxide : O2 − + H + → HOO• (20.16)


Electron scavenger : HOO + e − → HOO − (20.17)

Formation of H 2 O2 : HOO − + H + → H 2 O2 (20.18)

OH· + pollutant + O2− → products ( CO2 , H 2 O, etc. ) (20.19)

20.3.4  M
 esoporous Materials as Supports
for Photocatalytic Destruction

Mesoporous materials have attracted research interest owing to their high surface
area, high pore volume and high pore diameters, and hence it has been used as a
supporting material for several applications [44–47]. In the recent research concern,
Ti-containing mesoporous molecular sieves such as MCM-41, MCM-48, SBA-1,
SBA-15 and HMS have been utilized as active and selective photocatalysts for the
decomposition of NOx. Because such mesoporous materials are possessing high
surface area with high pore diameter, high thermal and hydrothermal stability which
are well-suitable features as catalyst/catalyst-supported, adsorbent and host-guest
encapsulation materials [48–50]. This new family of mesostructure with hierarchal
pore network catalyst is expected to demonstrate enhanced reactivity, selectivity
and resistance to deactivation. The isomorphous substitution of transition metal ions
such as Ti, Mo, Cu, V and Cr, into the framework of mesoporous materials has
remarkable catalytic and photocatalytic applications. The incorporation of transi-
tion metal atoms into the framework can create both acidic and redox sites which
make such materials potentially active photocatalytic materials for the photocata-
lytic decomposition of NOx. The schematic representation of framework of
MCM-41 and isomorphous substitution of metals is illustrated in Fig. 20.11.

 i-Containing Mesoporous Materials for Photocatalytic


T
Decomposition of NOx

Ti-MCM-41 and Ti-HMS have been synthesized and used as an efficient photocata-
lysts. The photocatalytic destruction of NO into N2 and O2 was examined over
Ti-MCM-41, HMS and TS-1 catalysts, and the photocatalytic performance with
respect of selectivity of N2 and N2O is illustrated in Fig. 20.12.
Generally, the photocatalytic destruction of NOx depends on the type of porous
materials which can be used as a support. From the figure, it is clearly evident that
Ti-MCM-41 material has showed higher selectivity of N2 than that of Ti-HMS and
TS-1 materials [51]. This higher activity and selectivity of Ti-MCM-41 is due to the
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 535

Fig. 20.11  Schematic illustration of framework of MCM-41 molecular sieves: (a) neutral, (b) M
(divalent metal ions) and (c) M (trivalent metal ions)

formation of well-dispersed Ti-oxide species within the framework and/or cavities


of Ti-MCM-41. Hence, the obtained results clearly indicated that isolated and tetra-
hedrally coordinated Ti-oxide species with well dispersion can act as active sites for
the photocatalytic destruction of NO into N2 and O2. The yield and selectivity of N2
and N2O in the direct photocatalytic destruction of NO over various amounts of
Ti-loaded Ti-MCM-41, Ti-HMS and Bulk TiO2 catalysts are listed in Table 20.1.
Furthermore, Hu et al. [51] have also investigated the mesoporous materials with
various Ti metal contents and characterized by using various in situ spectroscopic
techniques such as photoluminescence, FTIR, UV-Vis and XAFS spectroscopy.
They have demonstrated on the influence of the Ti contents in the Ti-MCM-41 on
the selectivity of N2 for the photocatalytic destruction of NO by using FTIR and
photoluminescence as shown in Fig. 20.13.
From Fig.  20.13, it can be observed that the total amount of the tetrahedral
Ti-oxide species exposed at the surface of Ti-MCM-41 can be monitored by the
intensity of the δasym NH3 IR band. On the other hand, the total amount of isolated
tetrahedrally coordinated Ti-oxide species can be seen from the photoluminescence
spectra of Ti-MCM-41. From the above results, it can be concluded that highly
536 K. Nanaji et al.

Fig. 20.12  Conversion of N2 and N2O for the photocatalytic destruction of NO over Ti-MCM-41,
Ti-HMS and TS-1 catalysts under UV irradiation at 295 K [51]

Table 20.1  The yield and selectivity of N2 and N2O in the direct photocatalytic destruction of NO
over various photocatalysts [51]
Yield (μmol/g of cat. h) Selectivity (%)
Catalysts Ti content in Wt.% N2 N2O Total N2 N2O
Ti-MCM-41 0.15 0.60 0.20 0.80 75 25
Ti-MCM-41 0.60 1.50 0.40 1.90 79 21
Ti-MCM-41 0.85 0.40 0.45 0.85 47 53
Ti-MCM-41 2.00 0.25 0.75 1.00 25 75
TS-1 0.60 0.85 0.70 1.55 55 45
Ti-HMS 0.60 0.95 0.45 1.40 68 32
Ti-HMS 1.26 16.0 3.0 19.0 84 16
Ti-HMS 2.30 13.0 4.5 17.5 74 26
Ti-HMS 11.6 7.5 6.0 13.5 56 44
Bulk – 3.5 9.0 12.5 28 72

dispersed and isolated tetrahedrally coordinated Ti-oxide species in the catalyst can
act as active sites for the photocatalytic destruction of NO into N2 and O2. Anpo and
co-workers [52] have prepared a bimetal substituted mesoporous V-Ti-MCM-41
catalyst by using a simple photo-assisted synthesis method, and they have noticed
an absorption spectrum with remarkable red shift compared with Ti-MCM-41. The
above-obtained results clearly suggested that bimetal substituted mesoporous V-Ti-­
MCM-41 catalyst can absorb light of longer wavelengths in the near-visible light
region. Furthermore, substitution of two different transition metal ions in the frame-
work of mesoporous photocatalysts can utilize the most efficient solar energy
resource.
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 537

Fig. 20.13  Photocatalytic destruction of NO into N2 and N2O: (a) infrared band of δasym NH3
adsorbed on Ti(IV5) and (b) photoluminescence spectra of Ti-MCM-41 with different Ti con-
tents [51]

CrS-1 CrS-1 N2 Selec,


N2 Selec, 45%
λ>450nm λ>270nm 97%
7

6 d Cr-HMS
Reaction Time/ h

λ>450nm
5 c
b Cr-HMS
4 λ>270nm
a
3
Light on
2

0
0.0 0.2 0.4 0.6 0.8 1.0
N2 Yields/mol. g-cat-1

Fig. 20.14  The photocatalytic destruction of NO by using Cr-HMS and CrS-1 under UV
(λ > 270 nm) as well as visible (λ > 450 nm) light irradiations [53]

Anpo et al. [53] have also reported that Cr-HMS also exhibited higher photocata-
lytic destruction of NO under both UV and visible light irradiation at 275 K. The
reaction time profiles of the photocatalytic destruction of NO on Cr-HMS and CrS-1
under both UV and visible light irradiation are shown in Fig. 20.14. It is observed
538 K. Nanaji et al.

from Fig. 20.14 that the yields of N2 increased with respect to irradiation time and
the selectivity for N2 (97%) under visible light irradiation showed higher value as
compared with UV light irradiation (45%). From the obtained results, it is con-
cluded that Cr-HMS can act as an active photocatalyst under both UV and visible
light irradiation.
The preparation and photocatalytic destruction of NO over Ti-HMS mesoporous
materials were investigated. It was found that Ti-HMS catalyst with highly dis-
persed isolated tetrahedral Ti species was found to be active species for high selec-
tivity of N2. Nevertheless, the bulk TiO2 and Ti-HMS (10) photocatalysts were
found to be favouring the formation of N2O due to the aggregated titanium oxide
species with octahedral coordination. The effects of Ti content and product distribu-
tion of the photocatalytic destruction of NO on various titanium oxide-based meso-
porous catalysts are represented in Table 20.1.

 hotocatalytic Oxidation of VOCs over Ti-Containing


P
Mesoporous-Based Photocatalysts

VOCs are very harmful and hazardous compounds, and hence several abatement
technologies of VOCs have been developed. However, there is no single method
available which focused on high energy consumption, low removal efficiency and
environmental friendliness. In this regard, the dual functionality of adsorption and
photocatalytic destruction of VOCs is one of the most promising techniques because
of the high removal efficiency. The design and developments of innovative TiO2-­
supported nanoporous composite photocatalysts are strongly required to enhance
the photocatalytic destruction of VOCs. The photocatalytic activity is mainly asso-
ciated with the amount of organic compounds adsorbed on the surface of the photo-
catalysts, and then the adsorbed organic compound is easily decomposed by the
photocatalysts under the solar irradiation.

20.4  M
 echanism and Activity of Titanium-Based
Photocatalysts for Photocatalytic Destruction of NOx
and VOCs

The photocatalytic active sites and their mechanisms of photocatalytic decomposi-


tion reactions are briefly described in the present section. Yamashita and Anpo [54]
have reported the relationship between the coordination number of Ti-oxide species
and the selectivity of N2 in the photocatalytic destruction of NO by using Ti-oxide
photocatalysts as shown in Fig. 20.15.
It can be observed from Fig. 20.15 that the selectivity of N2 is clearly depending
on the coordination number of the Ti-oxides located within the framework of zeo-
lites. From the obtained results, it is noted that a highly selective photocatalytic
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 539

Fig. 20.15  Correlation between the coordination number of Ti-oxides and the selectivity of N2 in
the photocatalytic destruction of NO using Ti-oxide photocatalysts [54]

destruction of NO into N2 and O2 can be achieved by using highly dispersed and


isolated tetrahedrally coordinated Ti-oxide species in the zeolite framework.
Nevertheless, the aggregated Ti-oxides with octahedral coordination showed the
predominant formation of N2O as product. The photocatalytic selectivity of N2 is
strongly depending on the nature of active sites and type of catalysts used in the
decomposition reaction. The reaction mechanism of the photocatalytic destruction
of NO into N2 and O2 on tetrahedrally coordinated Ti-oxide-supported zeolite cata-
lyst under UV light irradiation was proposed by Matsuoka and Anpo [55] as shown
in Fig. 20.16.
It was found that Ti-oxide species prepared within the framework of zeolites
have revealed a unique local structure as well as a high selectivity in the oxidation
of organic substances with H2O2. The previous studies clearly demonstrated that the
local structure and the dynamics of photochemical reactivities of the well-defined
transition metal oxides (titanium, chromium oxides, etc.) incorporated into the cavi-
ties and frameworks of various zeolites and zeo-type materials play a vital role in
the photocatalytic decomposition reaction.
Anpo et al. [56] have investigated the photocatalytic reaction of NO in the pres-
ence of various kinds of hydrocarbons such as methane, ethane and propylene. They
found that the photocatalytic efficiency of the NO strongly depends on the kind of
hydrocarbons used. They also proposed the important role of the intermediate spe-
cies formed from NO and hydrocarbon radicals such as the propyl radicals, which
were subsequently followed by further reaction with NO to produce N2, and then the
overall reaction of NO in the presence of hydrocarbons enhances the photocatalytic
efficiency.
540 K. Nanaji et al.

Fig. 20.16  Mechanism of the photocatalytic destruction of NO into N2 and O2 on tetrahedrally


coordinated Ti-oxide-supported zeolite catalyst under UV and visible light irradiation [54, 55]

20.5  Summary and Prospective

Photocatalytic destruction of air pollutants using semiconductor nanostructured


TiO2 photocatalyst has been extensively investigated as a potential and effective
technology for air purification because it can directly decompose the pollutants into
harmless counterparts under ambient atmosphere. Particularly, highly functional-
ized nanostructured titanium oxide-based photocatalysts can be a perfect candidate
for removing low concentration pollutants (even sub-ppm levels) in outdoor as well
as indoor environments. Among several proposed technologies, the photocatalytic
destruction of NOx/VOCs is recognized as a promising technology owing to the
excellent selectivity to N2 with considerable NOx conversion, harvesting based on
solar light, energy saving, low-cost technology, environmental energy and conver-
sion of NOx into nitrates (NO3−) which may recover as a conceivable raw material
for fertilizers.
It has been identified that the dual functional adsorbent/photocatalyst system
could be one of the effective and economic techniques for the removal of air pollut-
ants in the gaseous phase because the removal efficiency of pollutant over the com-
bination of adsorption/desorption with photocatalytic oxidation system was three
times higher than the normal photocatalytic oxidation system. Thus, these
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 541

adsorbent/photocatalyst composite materials deserve research attention due to their


combination properties of adsorption and photocatalysts which enhance the activity
and selectivity towards specific air pollutants. A combination of these fascinating
and unique features of the photocatalytic materials could provide a better approach
in the utilization of solar energy as the most abundant and safe energy source, and
these photocatalytic composite materials are promising candidates for the effective
reduction of toxic contaminants into harmless products at ambient temperature.

Acknowledgement The authors greatly acknowledge the funding support from Technical


Research Centre (TRC) project (Ref. No. AI/1/65/ARCI/2014 (c)) sponsored by the Department of
Science and Technology (DST), Government of India, New Delhi, India.

References

1. WHO. (2016). Ambient air pollution: A global assessment of exposure and burden of disease
(pp. 40–42).
2. Li, Y., Guan, D., Tao, S., Wang, X., & He, K. (2018). A review of air pollution impact on sub-
jective well-being: Survey versus visual psychophysics. Journal of Cleaner Production, 184,
959–968.
3. Brown, S. K., Sim, M. R., Abramson, M. J., & Gray, C. N. (1994). Concentrations of volatile
organic compounds in indoor air—A review. Indoor Air, 4, 123–134.
4. Weon, S., He, F., & Choi, W. (2019). Status and challenges in photocatalytic nanotechnology
for cleaning air polluted with volatile organic compounds: Visible light utilization and catalyst
deactivation. Environmental Science Nano, 6, 3185–3214.
5. Centi, G., Ciambelli, P., Perathoner, S., & Russo, P. (2002). Environmental catalysis: Trends
and outlook. Catalysis Today, 75, 3–15.
6. Lasek, J., Yu, Y.-H., & Wu, J.  C. S. (2013). Removal of NOx by photocatalytic processes.
Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 14, 29–52.
7. Shote, A. S., Betiku, E., & Asere, A. A. (2019). Characteristics of CO and NOx emissions from
combustion of transmethylated palm kernel oil-based biodiesel blends in a compression igni-
tion engine. Journal of King Saud University—Engineering Sciences, 31, 178–183.
8. Fujishima, A., Zhang, X., & Tryk, D. A. (2008). TiO2 photocatalysis and related surface phe-
nomena. Surface Science Reports, 63, 515–582.
9. Kumar, A., Kumar, K., & Krishnan, V. (2019). Sunlight driven methanol oxidation by aniso-
tropic plasmonic Au nanostructures supported on amorphous titania: Influence of morphology
on photocatalytic activity. Materials Letters, 245, 45–48.
10. Anpo, M. (Ed.). (1996). Surface photochemistry (pp. 1–186). West Sussex: Wiley.
11. Portela, R., & Hernández-Alonso, M. D. (2013). Environmental applications of photocatalysis.
Green Energy and Technology, 71, 35–66.
12. Lu, G., Linsebigler, A., & Yates, J. T. (1995). Photooxidation of CH3Cl on TiO2(110). A mech-
anism not involving H2O. The Journal of Physical Chemistry, 99, 7626–7631.
13. Anpo, M., Suzuki, E., & Giamello. (1990). New development in selective oxidation (p. 683).
Amsterdam: Elsevier.
14. Zhang, S. G., Ariyuki, M., Mishima, H., Higashimoto, S., Yamashita, H., & Anpo, M. (1998).
Photoluminescence property and photocatalytic reactivity of V-HMS mesoporous zeolites
effect of pore size of zeolites on photocatalytic reactivity. Microporous and Mesoporous
Materials, 21, 621–627.
15. Zhang, S. G., Higashimoto, S., Yamashita, H., & Anpo, M. (1998). Characterization of vana-
dium oxide/ZSM-5 zeolite catalysts prepared by the solid-state reaction and their p­ hotocatalytic
542 K. Nanaji et al.

reactivity: In situ photoluminescence, XAFS, ESR, FT-IR, and UV-vis investigations. The
Journal of Physical Chemistry. B, 102, 5590–5594.
16. Reddy, K. L., Kumar, S., Kumar, A., & Krishnan, V. (2019). Wide spectrum photocatalytic
activity in lanthanide-doped upconversion nanophosphors coated with porous TiO2 and Ag-Cu
bimetallic nanoparticles. Journal of Hazardous Materials, 367, 694–705.
17. Choi, W., Hong, S. J., Chang, Y. S., & Cho, Y. (2000). Photocatalytic degradation of polychlo-
rinated dibenzo-p-dioxins on TiO2 film under UV or solar light irradiation. Environmental
Science & Technology, 34, 4810–4815.
18. Kumar, A., Reddy, K. L., Kumar, S., Kumar, A., Sharma, V., & Krishnan, V. (2018). Rational
design and development of lanthanide-doped NaYF4@CdS-au-RGO as quaternary plas-
monic photocatalysts for harnessing visible-near-infrared broadband spectrum. ACS Applied
Materials & Interfaces, 18, 15565–15581.
19. Jan, Y. H., Lin, L. Y., Karthik, M., & Bai, H. (2009). Titanium dioxide/zeolite catalytic adsor-
bent for the removal of NO and acetone vapors. Journal of the Air & Waste Management
Association, 59, 1186–1193.
20. Tao, Y., Wu, C. Y., & Mazyck, D. W. (2006). Removal of methanol from pulp and paper mills
using combined activated carbon adsorption and photocatalytic regeneration. Chemosphere,
65, 35–42.
21. Tao, Y., Schwartz, S., Wu, C.-Y., & Mazyck, D. W. (2005). Development of a TiO2/AC com-
posite photocatalyst by dry impregnation for the treatment of methanol in humid air streams.
Industrial and Engineering Chemistry Research, 44, 7366–7372.
22. Binas, V., Venieri, D., Kotzias, D., & Kiriakidis, G. (2017). Modified TiO2 based photocata-
lysts for improved air and health quality. Journal of Materials, 3, 3–16.
23. Agrios, A.  G., Gray, K.  A., & Weitz, E. (2003). Photocatalytic transformation of

2,4,5-­trichlorophenol on TiO2 under sub-band-gap illumination. Langmuir, 19, 1402–1409.
24. Anpo, M. (2004). Preparation, characterization, and reactivities of highly functional titanium
oxide-based photocatalysts able to operate under UV–visible light irradiation: Approaches in
realizing high efficiency in the use of visible light. Bulletin of the Chemical Society of Japan,
77, 1427–1442.
25. Anpo, M., Aikawa, N., Kubokawa, Y., Che, M., Louis, C., & Giamello, E. (1985).

Photoluminescence and photocatalytic activity of highly dispersed titanium oxide anchored
onto porous Vycor glass. The Journal of Physical Chemistry, 89, 5017–5021.
26. Anpo, M., Aikawa, N., Kubokawa, Y., Che, M., Louis, C., & Giamello, E. (1985).

Photoformation and structure of oxygen anion radicals (O2-) and nitrogen-containing anion
radicals adsorbed on highly dispersed titanium oxide anchored onto porous Vycor glass. The
Journal of Physical Chemistry, 89, 5689–5694.
27. Yamashita, H., Honda, M., Harada, M., Ichihashi, Y., Anpo, M., Hirao, T., Itoh, N., & Iwamoto,
N. (1998). Preparation of titanium oxide photocatalysts anchored on porous silica glass by
a metal ion-implantation method and their photocatalytic reactivities for the degradation of
2-propanol diluted in water. The Journal of Physical Chemistry B, 102, 10707–10711.
28. Sen, S. E., Smith, S. M., & Sullivan, K. A. (1999). Organic transformations using zeolites and
zeotype materials. Tetrahedron, 55, 12657–12698.
29. Shioya, Y., Ikeue, K., Ogawa, M., & Anpo, M. (2003). Synthesis of transparent Ti-containing
mesoporous silica thin film materials and their unique photocatalytic activity for the reduction
of CO2 with H2O. Applied Catalysis A: General, 254, 251–259.
30. Yamashita, H., Ikeue, K., Takewaki, T., & Anpo, M. (2002). In situ XAFS studies on the effects
of the hydrophobic-hydrophilic properties of Ti-Beta zeolites in the photocatalytic reduction of
CO2 with H2O. Topics in Catalysis, 18, 95–100.
31. Takeuchi, M., Yamashita, H., Matsuoka, M., Anpo, M., Hirao, T., Itoh, N., & Iwamoto,
N. (2000). Photocatalytic decomposition of NO under visible light irradiation on the Cr-ion-­
implanted TiO2 thin film photocatalyst. Catalysis Letters, 67, 135–137.
20  Highly Functionalized Nanostructured Titanium Oxide-Based Photocatalysts… 543

32. Ikeue, K., Yamashita, H., Anpo, M., & Takewaki, T. (2001). Photocatalytic reduction of CO2
with H2O on Ti−β zeolite photocatalysts: Effect of the hydrophobic and hydrophilic proper-
ties. The Journal of Physical Chemistry B, 105, 8350–8355.
33. Anpo, M., & Che, M. (1999). Applications of photoluminescence techniques to the character-
ization of solid surfaces in relation to adsorption, catalysis, and photocatalysis. Advances in
Catalysis, 44, 119–257.
34. Anpo, M. (2000). Photofunctional zeolites: Synthesis, characterization, photocatalytic reac-
tions, light harvesting (pp. 1–236). New York: Nova Science Publishers Inc.
35. Yamashita, H., Ichihashi, Y., Anpo, M., Mitsuo, H., Louis, C., & Che, M. (1996). Photocatalytic
decomposition of NO at 275  K on titanium oxides included within Y-zeolite cavities: The
structure and role of the active sites. The Journal of Physical Chemistry, 100, 16041–16044.
36. Courbon, H., & Pichat, P. (1984). Room-temperature interaction of N18O with ultraviolet-­
illuminated titanium dioxide. Journal of the Chemical Society, Faraday Transactions, 80,
3175–3185.
37. Martra, G., Horikoshi, S., Anpo, M., Coluccia, S., & Hidaka, H. (2002). FTIR study of
adsorption and photodegradation of L-α-alanine on TiO2 powder. Research on Chemical
Intermediates, 28, 359–371.
38. Anpo, M., & Chiba, K. (1992). Photocatalytic reduction of CO2 on anchored titanium oxide
catalysts. Journal of Molecular Catalysis, 74, 207–212.
39. Hu, Y., Rakhmawaty, D., Matsuoka, M., Takeuchi, M., & Anpo, M. (2006). Synthesis, char-
acterization and photocatalytic reactivity of Ti-containing micro- and mesoporous materials.
Journal of Porous Materials, 13, 335–340.
40. Anpo, M., Yamashita, H., Matsuoka, M., Park, D., Shul, Y., & Park, S. (2000). Design and
development of titanium and vanadium oxide photocatalysts incorparated within zeolite cavi-
ties and their photocatalytic reactivities. Journal of Industrial and Engineering Chemistry,
6, 59–72.
41. Bowering, N., Walker, G.  S., & Harrison, P.  G. (2006). Photocatalytic decomposition and
reduction reactions of nitric oxide over Degussa P25. Applied Catalysis B: Environmental, 62,
208–216.
42. Cant, N., & Cole, J. (1992). Photocatalysis of the reaction between ammonia and nitric oxide
on TiO2 surfaces. Journal of Catalysis, 134, 317–330.
43. Shayegan, Z., Lee, C.-S., & Haghighat, F. (2018). TiO2 photocatalyst for removal of volatile
organic compounds in gas phase—A review. Chemical Engineering Journal, 334, 2408–2439.
44. Zhao, X.  S., Lu, G.  Q., & Millar, G.  J. (1996). Advances in mesoporous molecular sieve
MCM-41. Industrial and Engineering Chemistry Research, 35, 2075–2090.
45. Mercier, L., & Pinnavaia, T. I. (1997). Access in mesoporous materials: Advantages of a uni-
form pore structure in the design of a heavy metal ion adsorbent for environmental remedia-
tion. Advanced Materials, 9, 500–503.
46. Karthik, M., Vinu, A., Tripathi, A. K., Gupta, N. M., Palanichamy, M., & Murugesan, V. (2004).
Synthesis, characterization and catalytic performance of Mg and Co substituted mesoporous
aluminophosphates. Microporous and Mesoporous Materials, 70, 15–25.
47. Karthik, M., Tripathi, A.  K., Gupta, N.  M., Vinu, A., Hartmann, M., Palanichamy, M., &
Murugesan, V. (2004). Characterization of Co,Al-MCM-41 and its activity in the t-butylation
of phenol using isobutanol. Applied Catalysis A: General, 268, 139–149.
48. Beck, J.  S., Vartuli, J.  C., Roth, W.  J., Leonowicz, M.  E., Kresge, C.  T., Schmitt, K.  D.,
Chu, C. T. W., Olson, D. H., Sheppard, E. W., McCullen, S. B., Higgins, J. B., & Schlenker,
J. L. (1992). A new family of mesoporous molecular sieves prepared with liquid crystal tem-
plates. Journal of the American Chemical Society, 114, 10834–10843.
49. Corma, A. (1997). From microporous to mesoporous molecular sieve materials and their use
in catalysis. Chemical Reviews, 97, 2373–2420.
50. Raman, N. K., Anderson, M. T., & Brinker, C. J. (1996). Template-based approaches to the
preparation of amorphous, nanoporous silicas. Chemistry of Materials, 8, 1682–1701.
544 K. Nanaji et al.

51. Hu, Y., Martra, G., Zhang, J., Higashimoto, S., Coluccia, S., & Anpo, M. (2006). Characterization
of the local structures of Ti-MCM-41 and their photocatalytic reactivity for the decomposition
of NO into N2 and O2. The Journal of Physical Chemistry B, 110, 1680–1685.
52. Anpo, M., & Thomas, J. M. (2006). Single-site photocatalytic solids for the decomposition of
undesirable molecules. Chemical Communications, 3273.
53. Hu, Y., Wada, N., Matsuoka, M., & Anpo, M. (2004). Photo-assisted synthesis of V-MCM-41
under UV light irradiation. Catalysis Letters, 97, 49–52.
54. Yamashita, H., Yoshizawa, K., Ariyuki, M., Higashimoto, S., Anpo, M., & Che, M. (2001).
Photocatalytic reactions on chromium containing mesoporous silica molecular sieves
(Cr-HMS) under visible light irradiation: Decomposition of NO and partial oxidation of pro-
pane. Chemical Communications, 435–436.
55. Yamashita, H., & Anpo, M. (2003). Local structures and photocatalytic reactivities of the tita-
nium oxide and chromium oxide species incorporated within micro- and mesoporous zeolite
materials: XAFS and photoluminescence studies. Current Opinion in Solid State & Materials
Science, 7, 471–481.
56. Matsuoka, M., & Anpo, M. (2003). Local structures, excited states, and photocatalytic reac-
tivities of highly dispersed catalysts constructed within zeolites. Journal of Photochemistry
and Photobiology C: Photochemistry Reviews, 3, 225–252.
Chapter 21
Bandgap Engineering as a Potential Tool
for Quantum Efficiency Enhancement

Reddy Kunda Siri Kiran Janardhana, Raju Kumar, Tata Narsinga Rao,


and Srinivasan Anandan

21.1  Introduction

It is a well-known fact that pollution is increasing and humanity is under constant


exposure to harmful pollutants. Some studies even suggest the presence of an almost
equal level of pollutants in both indoor and outdoor environments. Since, an average
person spends almost 80% of his time in indoor environments [1], there is an
increase in the use of various daily comforting agents like air fresheners, deodor-
ants, paints, smoke-inducing agents, pesticides for pest control  which  combined
with household cooking leads to the formation of harmful by-products. These rea-
sons coupled with poor ventilation only exacerbates the situation. Such conditions
manifest themselves through deteriorated physical and mental health introducing a
number of symptoms which are collectively defined as sick building syndrome. This
leads to a reduction in human productivity, acute discomfort and in worst cases
chronic diseases [2].
To tackle this situation, there are some commonly used remedies like (1) dilu-
tion, where the indoor air is mixed with disinfected clean air; (2) use of filters like
high-efficiency particulate air (HEPA) filters at the openings of heating, ventilation
and air-conditioning (HVAC) systems; (3) UV irradiation, where high-energy UV
light is used to destroy DNA/RNA making pathogens inactive/ineffective; (4) instal-
lation of dehumidifiers with silica gels inside them, which enables them to trap vola-
tile organic compounds (VOC); and (5) anti-bacterial sprays (furthering the use of
comforting goods) which are usually alcohol based. Some of the advantages of such
methods include their almost zero contact with the running equipment, well-­
established protocols and most importantly a sense of quick remediation or the sen-
timent of instantly feeling good [3]. However, some of these solutions involve a lot

R. K. S. K. Janardhana · R. Kumar · T. N. Rao · S. Anandan (*)


Centre for Nano Materials, International Advanced Research Centre for Powder Metallurgy
and New Materials, Hyderabad, Telangana, India
e-mail: anandan@arci.res.in

© Springer Nature Switzerland AG 2021 545


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_21
546 R. K. S. K. Janardhana et al.

of initial manpower, very high installation costs, maintenance and a lot of power
input making them economically unviable. As a disruptively better solution, semi-
conductor photocatalysis is proving to address some of above-mentioned issues
quite efficiently. The major advantages of this technique are zero human interven-
tion, requiring only sunlight; rapid and economical end-to-end process; and almost
zero maintenance.

21.2  How Does Photocatalysis Work?

Photocatalysis is a technique in which (Fig. 21.1) sunlight acts as an accelerator to


a chemical reaction between encompassed reactants in the presence of semiconduc-
tors [4]. All semiconductors (SCs) are characterized by a forbidden region called the
bandgap in between the conduction band (CB) and valance band (VB). When the
energy of incident light is higher than the bandgap of the SC, the e−/h+ pairs present
in the SC separate, and the electrons get excited to the CB of the SC, whereas the
holes remain in the VB. These e−/h+ pairs participate in the reduction and oxidation
of pollutants, respectively. These e−/h+ pairs have an inherent nature of recombining
back, and thus, effective utilization of these e−/h+ is a must for quantum efficiency
(QE) enhancement. Sunlight, which is abundantly available has only less than 2%
UV light.  Thus, for effective utilization of the available spectrum, visible light-­
active photocatalysts are required [5]. TiO2 and ZnO are borderline UV-active mate-
rials, and hence it is easier to convert them to visible light-active materials that can
utilize the maximum spectrum of the sunlight for effective pollutant remediation.
Since, indoor environment has a very high presence of visible light, modifying these
wide bandgap materials to make them visible light-active photocatalysts is an ele-
gant solution to our problem in hand.

Fig. 21.1  Mechanism of photocatalysis


21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 547

Fig. 21.2  Band-edge positions of some SCs. Reprinted with permission from [6]. Copyright @
Elsevier 2004

The required bandgap is determined by the redox potential of the chemical entity


to be eliminated. For instance (Fig. 21.2), the VB potential of any SC must be higher
than 2.6 V (vs NHE, pH = 0) for the generation of •OH radicals which completely
mineralizes pollutant species. Similarly, the CB must be at −0.05  V (vs NHE,
pH = 0) so that molecular oxygen reacts with the photo-produced electrons. In addi-
tion, some overpotentials are needed which increase the bandgap requirement, and
thus a bandgap of more than 2.6  eV is required for real-time applications which
corresponds to ~480 nm or lesser wavelength in the blue light regime. People are
currently using many white LEDs in indoor environments which are composed of
red, blue and green wavelength regions. If not the sun, then the blue wavelength
region in the LED bulbs can excite photocatalysts in indoor environment. Therefore,
researchers are more attracted towards the designing of visible light-driven photo-
catalyst materials for indoor environment that can effectively work in low-intensity
indoor lighting, with minimal human intervention and at the same time cost-­effective
for practical uses [7].

21.3  Visible Light-Driven Photocatalysts

The conventional strategy of making visible light-active SCs is by doping with cat-
ionic or anionic entities, coupling with other SCs and plasmonic excitation. Though
generally sol-gel method is used, and numerous methods for doping are also avail-
able. Doping of cations (Fig. 21.3) like Ba, Co, Cu, Cr, Ni, Fe, V, Mn, Cd, Ce, Ta,
Ag, Bi, La, Nd, Mo, Nb [8–25] and many more into the host lattice is commonly
used for reducing bandgap and inducing visible light excitation. This doped entity
substitutes the host metal ion and introduces defect states near the CB which reduces
548 R. K. S. K. Janardhana et al.

Fig. 21.3  Formation of defect bands in SCs

the bandgap of the host material. Similarly, many other non-metallic elements like
B, N, C, S, P, etc. [26–30] can also be doped. These entities lead to the formation of
defect band near the VB. Although, the cationic and anionic defect bands lead to
visible light absorption, a trade-off with QE is seen, and these materials usually
have lower QE in visible light. At very low dopant concentrations, these defect
bands are useful by enhancing charge separation, but after a threshold limit, these
defect bands act as recombination centres by reducing mobilities acting as another
factor for QE decrement.
Electron and holes have an inherent tendency to lower their energy, and hence
electron moves towards more positive potentials and the holes towards more nega-
tive potentials (vs NHE, pH = 0). Taking advantage of this behaviour, the photocata-
lytic efficiency can be increased by coupling one photocatalyst (Fig.  21.4) with
another photocatalyst such that a small difference in the band-edge positions is pres-
ent between the photocatalysts. An example of such a system is the anatase and
rutile form of TiO2. Anatase has a bandgap of 3.2 eV and rutile has a bandgap of
3.0 eV. When these are coupled in such a way that the anatase/rutile is 80:20, we get
enhanced photocatalytic activity as the photo-excited electrons in the CB of anatase
can jump onto the CB of rutile increasing the e−/h+ carrier lifetimes by decreasing
recombination. Such careful arrangement of the VB and CB of many photocatalysts
(not only UV activated but also visible light activated) prevents e−/h+ recombination
increasing their lifetimes by facilitating the transfer of electrons and holes from one
SC to the other; however these strategies suffer from lower QE [31–34]. Similarly,
the free electrons present in the metals oscillate collectively when an incoming light
of similar energy is incident on these metals. Also known as surface plasma reso-
nance, it causes excessive charge build-up or higher charge density to be present
locally, and this favours electron transfer from the metal to the CB of the photocata-
lyst from where they are utilized for photocatalytic reactions. However, only some
metals like Au, Ag, Pt and Cu etc. [35–38] show this effect, and not all metals are
suitable for this effect to be noticed in the visible region. It is, therefore, understood
that SC photocatalysts need electrons with high reduction power and holes with
high oxidations power for effective and efficient photocatalytic activity. A brief
overview of the current trend in the surface modification of the SCs is the matter of
investigation for the current book chapter.
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 549

Fig. 21.4 Schematic
illustration of a coupled SC

21.3.1  Novel Strategy: Bandgap Engineering

The conventional approach for the selection of a SC is based on the positions of the
energy levels of VB and CB. For efficient degradation, deeper valance bands and
higher conduction bands are needed. In fact, one can increase the degradation activ-
ity based on the position of the VB and CB. Pt-grafted WO3 shows excellent oxida-
tion capability of acetaldehyde (pollutant) to harmless CO2 under visible light [39].
Since, the CB of WO3 (0.3–0.5 V vs SHE) is below the single-electron reduction of
oxygen (−0.067 V vs SHE), the excited electrons cannot reduce molecular oxygen.
However, the reduction potential is high enough for multi-electron reduction of
oxygen (O2  +  2H+  +  2e  →  H2O2, 0.68  V; or four-electron reduction,
O2  +  2H2O  +  4H+  +  4e  →  4H2O, 1.23  V), and it is this multi-electron reduction
(MER) that helps in the degradation of pollutants. Frei et al. [40] first reported the
Zr(IV)-O-Cu(I) bimetallic hetero-assemblies where the oxo-bridge acted as a pho-
toinduced redox centre. When excited, the oxo-bridged redox centre promoted the
electron transfer from the Zr(IV) to the Cu(I) forming Zr(III)-O-Cu(II) which in the
presence of CO2 reduced to form CO.  Nakamura et  al. [41] carefully selected
Ti(IV)-O-Ce(III) as a redox centre and reported that after excitation, the formation
of Ti(III)-O-Ce(IV), confirmed by X-ray absorption near-edge structure (XANES)
analysis, helps in the oxidative decomposition of 2-propanol into acetone and CO2.
In addition, the quantum efficiency (QE) was reported to be higher by almost 9
times for acetone and 4.75 times for CO2 formation than that of N-doped TiO2.
Since, the mononuclear Ti(IV) or Ce(III) samples did not catalyse any 2-propanol
conversion, it was thus hypothesized that the metal-to-metal charge transfer
(MMCT) induced by the visible light was responsible for the photocatalytic activity.
Based on MMCT, Irie et  al. [42] proposed that the same phenomena, i.e. charge
transfer, can be observed from atomic metal ions to the CB of TiO2 which are made
up of 3d orbitals, and to prove it, Cr(III) was grafted on TiO2 as a Cr(IV)/Cr(III)
550 R. K. S. K. Janardhana et al.

redox couple which has a E0 of 2.1 V. Rather than the VB electron getting excited,


the electron from the Cr(IV)/Cr(III) redox couple can directly be injected to the
CB. Thus, the Cr(IV)/Cr(III) sites acted as the oxidation sites oxidizing pollutants,
whereas the electron injected to the CB acted in reducing the molecular oxygen,
thereby getting consumed. This MMCT is similar to interfacial charge transfer
(IFCT) which was formulated by Hush et al. and Creutz et al. [43–45]. The same
IFCT phenomenon coupled with MER was later used by Irie et al. to transfer excited
VB electrons directly to the surface-grafted metal ions under visible light and
reported Cu2+-grafted TiO2 and WO3 [46]. When visible light falls on pristine TiO2,
the large bandgap prohibits any absorption and thus exhibits no catalytic activity.
But when Cu2+ is grafted on TiO2, the excited valance band electrons are directly
injected to the surface Cu2+ ions and lead to the formation of Cu1+. Since, Cu1+ is
unstable, it again goes back to its former state by transferring the electron to the
oxygen. Similarly, to explain the high photocatalytic activity of Cu2+-grafted WO3,
it was reported that MER also takes place since the conduction band potential of
WO3 is not enough to reduce oxygen through single-electron reduction. The current
discussed strategies (Fig. 21.5) help in increasing the QE for conventional doping
strategies from 3.9% of N-doped TiO2 to 8.8% of Cu2+-TiO2 and 17.5% of Cu2+-WO3.
Similarly, loading elements like Fe and Ag have also been found to be a suitable
candidate which exhibits the IFCT in addition to Cu2+, Cr3+ and Ce3+ which are
explained in the following section. The Fe3+/Fe2+ reduction potential (0.771 vs SHE,
pH = 0) is much higher than required for MER and follows a pattern which is simi-
lar to copper, where the Fe2+ gets converted back to Fe3+, and since it is having very
high positive potential, it was expected to be active in the visible light (~500 nm
range) [47]. Even Ag+/Ag has a reduction potential that is almost equal to that of
Fe3+/Fe2+, and after excitation, the valance band electrons are directly transferred to
the surface-grafted metal cations by IFCT, where electrons are used for multi-­
electron reduction (MER) [48]. Unlike doping, grafting entities on the surface does
not lead to any structural changes. Surface modification has no influence on

Fig. 21.5  Various excitation mechanisms in (a) N-doped TiO2, (b) Cu2+-grafted TiO2 and (c) Cu2+-
grafted WO3 [46, 47]
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 551

bandgap, but it significantly improves the charge separation by reducing the charge
recombination (unlike doping) process which enhances the photocatalytic ability of
the catalyst.
As a follow-up to the existing methods, Kumar et al. [49] (Figs. 21.6 and 21.7)
reported a Cu2+-modified ZnO photocatalyst. Prior, ZnO nanomaterials were syn-
thesized by flame spray pyrolysis (FSP) followed by Cu2+ modification. Notably,
surface-modified SCs showed no structural changes, indicating that no defect bands
were formed in the SC after Cu2+ surface modification. The surface modification
enables charge separation, in which electrons have high reduction potential and the
holes have high oxidation power, which can lead to efficient degradation of pollut-
ants. A similar observation was reported by Yin et al. [50], but under UV light, the
presence of Cu2+ grafting on the surface of Nb3O8− enhanced the production of
CO. Under light irradiation, photogenerated electrons are excited to the CBs and
moved to Cu2+ nanoclusters which act as electron sink and get converted to Cu1+.
Since, Cu1+ is highly unstable, electrons are transferred to reduce CO2 to form
CO.  On the other hand, generated holes were used for water oxidation to form
molecular oxygen. Shoji et al. [51] also reported a CuxO-modified SrTiO3 for pho-
tocatalytic CO2 reduction reactions. It was reported that the mix of the valance states
of Cu was crucial for the CO2 reduction process. Though the QE was not reported,
the CO2 utilization of CuxO-modified SrTiO3 was reported to have been almost dou-
ble than that of pure SrTiO3 under the same conditions.
Kumar et al. [52] reported a Fe3+-modified ZnO (Fig. 21.8) which was used for
killing E. coli pathogen under the visible light illumination. The Fe3+ ion-induced
IFCT in the visible light coupled with the action of the surface-grafted Fe3+ as a co-­
catalyst promoted multi-electron reduction effectively and improved the photocata-
lytic activity of these photocatalysts. The reduction reactions with oxygen give rise
to many reactive oxygen species (ROS) which helped in the oxidative destruction of
the E. coli cell wall. Another point to be noted is that living microbial entities can
negate the effect of these ROS because of the presence of antioxidants secreted by
them. Thus, a low concentration of ROS may not cause significant damage, but
higher concentration of these ROS can cause damage. Since, the E. coli bacteria
were completely destroyed under the visible light illumination, it can be inferred
that there was a high concentration of ROS present and thus these surface-modified
photocatalysts show exceptional activity.
Liu et al. [53] and Qiu et al. [54] also reported a Cu2+ ion-modified TiO2 where
Cu was grafted on the TiO2 by a simple impregnation technique. The uniqueness
2+

of this method is that unlike the previous reported SCs, the current reported SC can
function in the dark as an anti-bacterial agent. During the normal visible light illu-
mination on catalyst, excited electrons are transferred to the surface Cu2+ by IFCT
to form Cu1+ species. Since, Cu1+ is highly unstable and to get the stable form Cu2+,
electrons are released that react with the outer cell wall of pathogens effectively
destroying them. It was also reported that the presence of electrons in the Cu2+ states
is more preferred due to their higher positive potential in comparison to the CB of
552 R. K. S. K. Janardhana et al.

Fig. 21.6 (a, b) SEM images, (c, d) TEM images and (e, f) UV DRS studies of ZnO nano-spheres
and ZnO nanorods synthesized by flame spray pyrolysis, respectively. Reprinted with permission
from [49]. Copyright @ American Chemical Society 2014

TiO2. A small possibility exists that some of amount of these electrons can be
trapped at the interface of TiO2 and Cu2+ because of the different bond length at the
surface creating defects in the bandgap leading to a small increase in the electron-­
hole pair lifetime [53, 54].
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 553

Fig. 21.7  Schematic mechanism of the inactivation of E. coli bacteria. Reprinted with permission
from [49]. Copyright @ American Chemical Society 2014

Fig. 21.8  Visible light activity of (a) control sample, (b) ZnO nanorod, (c) Fe-modified ZnO
nanorod, (d) ZnO nano-sphere and (e) Fe-modified ZnO nano-sphere [52]

21.3.2  Surface Grafting of Doped Semiconductors

Surface grafting with Cu2+, Fe3+ and Ag1+ is a promising concept, but the IFCT phe-
nomena is limited only to the surface and metal oxide nanocluster interface. As
described before, doping leads to bandgap reduction and induces visible light acti-
vation. But unintentional high doping forms narrow bandgap which increases
recombination centres and results in lower photocatalytic ability. For oxidation
reaction like VOC degradation, doping with anions leads to the formation of defect
bands above the VB and reduces the QE of the photocatalyst; it is understood that
modifying the VB would be an unfavourable choice, and it is of the best interest to
not modify the VB of the SC. Thus, only the CB can be modified to induce visible
554 R. K. S. K. Janardhana et al.

light absorption where the cationic dopants promote the absorption of visible light
by forming isolated defect states below the CB and also help in  narrowing the
bandgap.
Maintaining a strict conduction that VB should remain unchanged, Yu et al. [55]
have developed a Ga3+- and W6+-doped Ti1−3xWxGa2xO2 where it was anticipated that
the W 5d bands contribute to the VB of TiO2 which are composed mostly of 3d
orbitals. Other requirements like charge neutrality and similar ionic radii  were
almost met for the above two doped entities. Doping of the entities undoubtedly led
to enhanced absorption in the visible region of the spectrum, but the QE of the
Ti1−3xWxGa2xO2 was very low (<0.1%), even less than N-doped TiO2. However, after
the surface modification with Cu2+, the QE remarkably increased to 13% which is
significantly higher than pristine TiO2 or Cu2+-modified TiO2. The doped impurity
bands impede charge mobility and also assist in recombination process, leading to
lower QE, but after Cu2+ modification, the surface-grafted ions effectively aid in the
extraction of the electron from defect states of CB and also through IFCT, leading
to enhanced photocatalytic behaviour.
Similarly, Qui et al. [56] also reported a Cu2+ ion-modified Mo and Na co-doped
SrTiO3 which showed visible light activity. The doped entities help in the formation
of defect bands below the CB of SrTiO3 and assist absorption of the visible light,
whereas the surface-grafted Cu2+ entities assist in effectively extracting electrons.
Following the same concept, Anandan et al. [57] synthesized a visible light-active
Ce-doped ZnO photocatalyst. A series of increasing amount of Ce was doped into
the ZnO followed by Cu2+ grafting. ZnO has a bandgap of ~3.4 eV and is not active
in visible light illumination but active in UV light. From the diffused reflectance
spectroscopy (DRS) analysis, pristine ZnO has absorption starting from 390  nm.
Doping with Ce introduced impurity state below the CB which caused a red shift of
the excitation wavelength and strong absorption in visible light. The authors
extended the IFCT concept and synthesized a Cu2+-grafted ZnO, and the UV-Vis
DRS measurements indicate an absorption for Cu2+-grafted ZnO in the visible light.
The degradation of acetaldehyde was conducted under visible light illumination,
and the photocatalytic tests are of the following order: ZnO < Ce-doped ZnO < Cu2+-
modified ZnO < Cu2+-modified Ce-doped ZnO.
Owing to its wide bandgap, pristine ZnO has no photocatalytic activity under
visible light irradiation and cannot degrade acetaldehyde. Ce-doped ZnO has visible
light activity, but the defect states act as recombination centres, and hence marginal
improvement is observed. However, the Cu2+-modified ZnO has improved the acet-
aldehyde degradation capability, owing to the charge separation by IFCT of surface-­
grafted Cu2+ under the visible light. Moreover, a Cu2+-grafted Cd-doped ZnO [58]
visible light-active SC was also reported by Anandan et al., and the results demon-
strated that Cu2+ grafting on the ZnO remarkably increased the photocatalytic
activity.
On the other hand, high-temperature calcination of TiO2 creates Ti3+ defect states
below the CB, owing to the oxygen vacancies. By taking advantage of this impurity
state, Liu et al. [60] prepared Cu2+ ion-modified Ti3+ self-doped TiO2 material, in
which Ti3+ impurity state below the CB caused bandgap narrowing and enhanced
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 555

the visible light absorption and the surface-grafted Cu2+ ions assisted in charge sep-
aration process, thereby maintaining a very high QE by IFCT and MER reactions.
Similarly, Kumar et al. [59] (Figs. 21.9, 21.10 and 21.11) reported a similar photo-
catalyst albeit with a different synthesis procedure using solution precursor plasma
spray (SPPS) process, in which the precursor for the TiO2 and Cu2+ is directly
injected in the plasma spray to produce Cu2+ ion-modified Ti3+ self-doped TiO2
material. This is a single-step in situ synthesis process, which reduces chemical
exposure, requires very less time and can be used for the mass production to com-
pete with the market requirements.

21.3.3  Energy-Level Matching as a Possible Alternative

It is clear that doping enhances visible light absorption, but at the same time, a clear
trade-off is seen in its QE. On the contrary, the surface grafting leads to an enhanced
QE, but its reported absorption is very low as the IFCT is only restricted to the sur-
face of the SCs. Liu et al. [61] reported a Fe-doped Fe-modified TiO2, where doping
increases the visible light absorption and at the same time the surface Fe-grafted
nanoclusters help in maintaining a very high QE. Though this surface grafting model
is already discussed in Sect. 21.3.2, one must look into the rationale behind this
selection. The defect bands of Fe form at approx. 0.3–0.5 eV lower than the CB of
TiO2. This defect band helps in absorbing visible light, giving rise to higher visible
light absorption. The excited VB electrons jump to the defect bands whose position
is near to the redox potential of Fe3+/Fe2+ (0.771 V vs SHE, pH = 0) and it is expected
that the transfer of the trapped electrons to the surface Fe3+/Fe2+ states is more effec-
tive on account of the almost similar energy levels. With such an approach, the QE
for 2-propanol degradation to CO2 in the visible light region reached about 47%. A
similar concept was utilized by Liu et  al. [62], who reported a Cu2+-modified
Nb-doped TiO2. Nb doping introduces defect band at the same energy level as the

Fig. 21.9 (a) Bandgap of self-modified TiO2 where the reduction cannot proceed via single elec-
tron and (b) Cu ion-modified self-doped TiO2 where the grafted co-catalyst acts as a centre for
multi-electron reduction reactions. Reprinted with permission from [59]. Copyright @ American
Chemical Society 2016
556 R. K. S. K. Janardhana et al.

Fig. 21.10  Schematic representation of the application of the in situ Cu2+ ion-modified Ti3+ self-­
doped TiO2. Reprinted with permission from [59]. Copyright @ American Chemical Society 2016

reduction potential of Cu2+/Cu1+. The defect band helps in boosting visible light
absorption, whereas the surface-grafted Cu2+ acts as an efficient electron collector,
enhancing oxygen reduction by multi-electron reduction reaction. The QE of these
reported photocatalysts was 25%. This is lower in comparison to Fe-doped
Fe-Modified TiO2 as the redox potential of Fe is far higher leading to absorption in
the sub-500 nm wavelength.
Similarly, Nanaji et al. [48] (Figs. 21.12 and 21.13) also reported a visible light-­
active Ag-modified Ag-doped TiO2 for various applications. Pure TiO2 has an insig-
nificant visible light photocatalytic activity as the bandgap of pure TiO2 is large for
visible light excitement. However, discrete and deep impurity bands are introduced
when Ag doping takes place, which starts at approximately 0.7 eV below the CB,
and upon visible light irradiation, the VB electrons get photo-excited to the impurity
states below the CB.  The position of these discrete impurity bands prohibits the
single-electron reduction of O2 to form O2−. Additionally, the carrier mobility is also
reduced which limits the photocatalytic activity. In stark contrast to these observa-
tions, when visible light was irradiated on Ag-grafted TiO2, the excited valence band
electrons are transferred to the surface silver ions due to IFCT. This electron transfer
aids in complete degradation of pollutants by forming reactive oxygen species
(ROS). The holes in the VB have sufficiently high potential to form •OH which oxi-
dize harmful acetaldehyde into harmless CO2 and degrade methylene blue (MB). In
the case of the Ag-doped and Ag-modified system, the Ag-doped impurity states
below the CB (0.7 eV below CB) and the reduction potential of Ag+/Ag (0.779 V)
are almost similar. This similarity boosts the transfer of the defect trapped electron
effectively to the surface Ag ions which further undergo multi-electron reduction
(i.e. two-electron reduction, O2  +  2H+  +  2e−  →  H2O2, 0.68  V; or four-electron
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 557

Fig. 21.11 (a) Ti3+ self-doped TiO2, (b) Cu ion-modified Ti3+ self-doped TiO2, (c) TEM images Cu
ion-modified Ti3+ self-doped TiO2 where the arrows and semi-circles indicate the surface-grafted
Cu2+ and (d) UV-Vis DRS plots. Reprinted with permission from [59]. Copyright @ American
Chemical Society 2016

reduction, O2 + 2H2O + 4H+ + 4e− → 4H2O, 1.23 V, etc.). This kind of approach of


energy-level matching boosts photocatalytic activity and can be used in indoor light
effectively.
Another approach was adopted by Liu et al. with a rationale that the QE of the
grafted SCs is sufficiently high (QE > 20%) with the efficient utilization of electrons
in the CB, and thus it was the oxidation activity of the holes that was limiting the QE
from the further increment. As Ti(IV)-O bond length is different in bulk and surface
sites, it promoted the grafted Ti(IV) nanocluster to act as a potential hole extractor/
acceptor and Cu(II) or Fe(III) as electron acceptor. During visible light illumination,
the photo-excited electrons are effectively transferred to the surface-grafted Cu(II)
or Fe(III) ions on the account of IFCT, and the holes are transferred to the grafted
Ti(IV). Such a scheme effectively enhanced separation between the electrons and
holes and this arrangement realizes a QE of 90% for Cu(II)-grafted Ti(IV)-TiO2 and
a QE of 92% for Fe(III)-grafted Ti(IV)-TiO2. These quantum efficiencies exceed
that of TiO2 under UV light illumination, which is one of the best-known photocata-
lysts [63, 64] (Table 21.1).
558 R. K. S. K. Janardhana et al.

Fig. 21.12 (a, b) Transmission electron microscopy images, (c) selected area electron diffraction
pattern and (d) lattice fringes of 101 planes of the Ag-modified-Ag-doped TiO2. Reprinted with
permission from [48]. Copyright @ Elsevier 2019

21.4  Conclusion and Future Works

In this chapter, the necessity of visible light-active catalysts, the concepts of grafting
and energy-level matching as potential bandgap engineering methods have been
discussed. Grafting enhances the visible light photocatalytic ability of the SC on
which the entities are grafted. The high QE is the result of visible light absorption
in the sub-500 nm region of the spectrum, depending on the ions grafted. Similarly,
the simultaneous doped and grafted SCs based on the concept of energy-level
matching also show enhanced visible light absorption and show reasonably high
QE. The holes have very high oxidation power since there is no introduction of any
defect bands near the VB, and the electrons in the CB are effectively used for oxy-
gen reduction reactions by MER.  A more effective utilization of these holes can
lead to an exceptionally high QE of 92% as seen in some of the cases and that these
photocatalysts show very good photocatalytic capabilities.
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 559

Fig. 21.13  Visible light excitation in (a) pure TiO2, (b) Ag-doped TiO2, (c) Ag-grafted TiO2,
and  (d) Ag-doped-Ag-grafted TiO2. Similarly, the indoor light degradation activity of (a) pure
methylene blue, (b) pure TiO2 and (c) silver-doped and silver-modified TiO2. Reprinted with per-
mission from [48]. Copyright @ Elsevier 2019

Table 21.1  Quantum efficiencies of photocatalysts that are discussed in this chapter (IPA stands
for isopropyl alcohol)
Sr. No. Material Photocatalytic activity QE(%) Refs.
1 N-doped TiO2 IPA degradation 3.9 [47]
2 Cu(II)-TiO2 IPA degradation 8.8 [46]
3 Fe(III)-TiO2 IPA degradation 22.0 [47]
4 Cr(III)-TiO2 IPA degradation 1.7 [42]
5 Cu(II)-WO3 IPA degradation 17.5 [46]
6 Cu(II)-(Sr1-yNay)(Ti1-xMox)O3 IPA degradation 14.5 [56]
7 Cu(II)-Ti1-3xWxGa2xO2 IPA degradation ~0.1 [55]
8 Nb-doped TiO2 IPA degradation 2.3 [62]
9 Cu(II)-Nb-doped TiO2 IPA degradation 25.3 [62]
10 Fe-doped TiO2 IPA degradation 6.5 [61]
11 Fe(III) Fe-doped TiO2 IPA degradation 47.3 [61]
12 Cu(II)-Ti(IV)-WO3 IPA degradation 21.2 [64]
13 Cu(II)-Ti(IV)-TiO2 IPA degradation 89.6 [63]
14 Fe(III)-Ti(IV)-TiO2 IPA degradation 92.2 [63]

A direction where this research can progress immensely is in the synthesis pro-
tocol. Currently, almost all methods use multiple steps for fabricating these photo-
catalysts. If newer methods which can fabricate the desired photocatalyst in a single
step are designed, it will bolster the use of these materials immensely as an eco-
nomically viable option. Also, sensitizing people about the adoption of these mate-
rials as a potential anti-fungal and anti-microbial agent must be encouraged and
560 R. K. S. K. Janardhana et al.

public exposure to such methods will be immensely helpful in adopting this tech-
nology. Additionally, common utilities like public washrooms and hospitals where
there is always a risk of high pathogenic load must be encouraged to adopt this
technology.

References

1. Srivastava, P. K., Pandit, G. G., Sharma, S., & Mohan Rao, A. M. (2000). Volatile organic com-
pounds in indoor environments in Mumbai, India. Science of the Total Environment, 255(1–3),
161–168.
2. Kostiainen, R. (1995, January). Volatile organic compounds in the indoor air of normal and
sick houses. Atmospheric Environment [Internet], 29(6), 693–702. Retrieved from https://link-
inghub.elsevier.com/retrieve/pii/1352231094003099.
3. Bolashikov, Z.  D., & Melikov, A.  K. (2009, July). Methods for air cleaning and protection
of building occupants from airborne pathogens. Building and Environment [Internet], 44(7),
1378–1385. https://doi.org/10.1016/j.buildenv.2008.09.001.
4. Fujishima, A., Rao, T. N., Tryk, D. A. (2000, June). Titanium dioxide photocatalysis. Journal
of Photochemistry and Photobiology C: Photochemistry Reviews [Internet], 1(1), 1–21.
Retrieved from https://linkinghub.elsevier.com/retrieve/pii/S1389556700000022.
5. Abe, R. (2010). Recent progress on photocatalytic and photoelectrochemical water splitting
under visible light irradiation. Journal of Photochemistry and Photobiology C: Photochemistry
Reviews [Internet], 11(4), 179–209. https://doi.org/10.1016/j.jphotochemrev.2011.02.003.
6. Carp, O. (2004). Photoinduced reactivity of titanium dioxide. Progress in Solid State Chemistry
[Internet], 32(1–2), 33–177. Retrieved from https://linkinghub.elsevier.com/retrieve/pii/
S0079678604000123.
7. Miyauchi, M., Irie, H., Liu, M., Qiu, X., Yu, H., Sunada, K., et al. (2016). Visible-light-sensitive
photocatalysts: Nanocluster-grafted titanium dioxide for indoor environmental remediation.
Journal of Physical Chemistry Letters, 7(1), 75–84.
8. Nithya, N., Bhoopathi, G., Magesh, G., & Kumar, C. D. N. (2018, December). Neodymium
doped TiO2 nanoparticles by sol-gel method for antibacterial and photocatalytic activ-
ity. Materials Science in Semiconductor Processing [Internet], 83, 70–82. https://doi.
org/10.1016/j.mssp.2018.04.011.
9. Zhu, X., Pei, L., Zhu, R., Jiao, Y., Tang, R., & Feng, W. (2018). Preparation and characteriza-
tion of Sn/La co-doped TiO2 nanomaterials and their phase transformation and photocatalytic
activity. Scientific Reports [Internet], 8(1), 1–14. https://doi.org/10.1038/s41598-­018-­30050-­3.
10. Mostoni, S., Pifferi, V., Falciola, L., Meroni, D., Pargoletti, E., Davoli, E., et  al. (2017).
Tailored routes for home-made Bi-doped ZnO nanoparticles. Photocatalytic performances
towards o-toluidine, a toxic water pollutant. Journal of Photochemistry and Photobiology A:
Chemistry [Internet], 332, 534–545. https://doi.org/10.1016/j.jphotochem.2016.10.003.
11. Yousefi, H. R., & Hashemi, B. (2019). Photocatalytic properties of Ag@Ag-doped ZnO core-­
shell nanocomposite. Journal of Photochemistry and Photobiology A: Chemistry, 375, 71–76.
12. Singh, N., Prakash, J., Misra, M., Sharma, A., & Gupta, R. K. (2017). Dual functional Ta-doped
electrospun TiO2 nanofibers with enhanced photocatalysis and SERS detection for organic
compounds. ACS Applied Materials & Interfaces, 9(34), 28495–28507.
13. Ahmad, M., Ahmed, E., Zafar, F., Khalid, N.  R., Niaz, N.  A., Hafeez, A., et  al. (2015).
Enhanced photocatalytic activity of Ce-doped ZnO nanopowders synthesized by c­ ombustion
method. Journal of Rare Earths [Internet], 33(3), 255–262. https://doi.org/10.1016/
S1002-­0721(14)60412-­9.
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 561

14. Andronic, L., Enesca, A., Vladuta, C., & Duta, A. (2009). Photocatalytic activity of cadmium
doped TiO2 films for photocatalytic degradation of dyes. Chemical Engineering Journal,
152(1), 64–71.
15. Baylan, E., & Altintas, Y.  O. (2019, July). Highly efficient photocatalytic activity of stable
manganese-doped zinc oxide (Mn:ZnO) nanofibers via electrospinning method. Materials
Science in Semiconductor Processing, 103.
16. Wang, T., Shen, D., Xu, T., & Jiang, R. (2017). Photocatalytic degradation properties of
V-doped TiO2 to automobile exhaust. Science of the Total Environment [Internet], 586,
347–354. https://doi.org/10.1016/j.scitotenv.2017.02.021.
17. Liu, L., Liu, Z., Yang, Y., Geng, M., Zou, Y., Shahzad, M. B., et al. (2018). Photocatalytic prop-
erties of Fe-doped ZnO electrospun nanofibers. Ceramics International [Internet], 44(16),
19998–20005. https://doi.org/10.1016/j.ceramint.2018.07.268.
18. Bai, J. Q., Wen, W., & Wu, J. M. (2016). Facile synthesis of Ni-doped TiO2 ultrathin nanobelt
arrays with enhanced photocatalytic performance. CrystEngComm, 18(10), 1847–1853.
19. Chang, C. J., Yang, T. L., & Weng, Y. C. (2014). Synthesis and characterization of Cr-doped
ZnO nanorod-array photocatalysts with improved activity. Journal of Solid State Chemistry
[Internet], 214, 101–107. https://doi.org/10.1016/j.jssc.2013.09.039.
20. Mittal, M., Sharma, M., & Pandey, O.  P. (2014). UV-Visible light induced photocatalytic
studies of Cu doped ZnO nanoparticles prepared by co-precipitation method. Solar Energy
[Internet], 110, 386–397. https://doi.org/10.1016/j.solener.2014.09.026.
21. AltIn, I., Sökmen, M., & Biyiklioʇlu, Z. (2016). Sol gel synthesis of cobalt doped TiO2 and
its dye sensitization for efficient pollutant removal. Materials Science in Semiconductor
Processing, 45, 36–44.
22. Jayakrishnan, A. R., Alex, K. V., Tharakan, A. T., Kamakshi, K., Silva, J. P. B., Prasad, M. S.,
et al. (2020). Barium-doped zinc oxide thin films as highly efficient and reusable photocata-
lysts. ChemistrySelect, 5(9), 2824–2834.
23. Jafari, H., Sadeghzadeh, S., Rabbani, M., & Rahimi, R. (2018). Effect of Nb on the structural,
optical and photocatalytic properties of Al-doped ZnO thin films fabricated by the sol-gel
method. Ceramics International [Internet], 44(16), 20170–20177. https://doi.org/10.1016/j.
ceramint.2018.07.311.
24. Wang, S., Bai, L. N., Sun, H. M., Jiang, Q., & Lian, J. S. (2013). Structure and photocatalytic
property of Mo-doped TiO2 nanoparticles. Powder Technology [Internet], 244, 9–15. https://
doi.org/10.1016/j.powtec.2013.03.054.
25. Sanchez Rayes, R. M., Kumar, Y., Cortes-Jácome, M. A., Toledo Antonio, J. A., Mathew, X.,
& Mathews, N. R. (2017). Effect of Eu doping on the physical, photoluminescence, and photo-
catalytic characteristics of ZnO thin films grown by sol–gel method. Physica Status Solidi (A)
Applications and Materials Science, 214(12), 1–9.
26. Wang, W.  K., Chen, J.  J., Gao, M., Huang, Y.  X., Zhang, X., & Yu, H.  Q. (2016).

Photocatalytic degradation of atrazine by boron-doped TiO2 with a tunable rutile/anatase
ratio. Applied Catalysis B: Environmental [Internet], 195, 69–76. https://doi.org/10.1016/j.
apcatb.2016.05.009.
27. Sakthivel, S., & Kisch, H. (2003). Daylight photocatalysis by carbon-modified titanium diox-
ide. Angewandte Chemie International Edition, 42(40), 4908–4911.
28. Asahi, R., Morikawa, T., Ohwaki, T., Aoki, K., & Taga, Y. (2001). Visible-light photocatalysis
in nitrogen-doped titanium oxides. Science, 293(5528), 269–271.
29. Yang, K., Dai, Y., & Huang, B. (2007). Understanding photocatalytic activity of S- and
P-doped TiO2 under visible light from first-principles. Journal of Physical Chemistry C,
111(51), 18985–18994.
30. Devi, L.  G., & Kavitha, R. (2014). Enhanced photocatalytic activity of sulfur doped TiO2
for the decomposition of phenol: A new insight into the bulk and surface modification.
Materials Chemistry and Physics [Internet], 143(3), 1300–1308. https://doi.org/10.1016/j.
matchemphys.2013.11.038.
562 R. K. S. K. Janardhana et al.

31. Di Paola, A., Palmisano, L., Venezia, A. M., & Augugliaro, V. (1999, September). Coupled
semiconductor systems for photocatalysis. Preparation and characterization of polycrystal-
line mixed WO 3/WS 2 powders. The Journal of Physical Chemistry B [Internet], 103(39),
8236–8244. Retrieved from https://pubs.acs.org/doi/10.1021/jp9911797.
32. Rawal, S. B., Bera, S., Lee, D., Jang, D.-J., & Lee, W. I. (2013). Design of visible-light pho-
tocatalysts by coupling of narrow bandgap semiconductors and TiO2: Effect of their rela-
tive energy band positions on the photocatalytic efficiency. Catalysis Science & Technology
[Internet], 3(7), 1822. Retrieved from http://xlink.rsc.org/?DOI=c3cy00004d.
33. Wu, L., Yu, J. C., & Fu, X. (2006). Characterization and photocatalytic mechanism of nano-
sized CdS coupled TiO2 nanocrystals under visible light irradiation. Journal of Molecular
Catalysis A: Chemical, 244(1–2), 25–32.
34. Luo, Z., Poyraz, A. S., Kuo, C. H., Miao, R., Meng, Y., Chen, S. Y., et al. (2015). Crystalline
mixed phase (anatase/rutile) mesoporous titanium dioxides for visible light photocatalytic
activity. Chemistry of Materials, 27(1), 6–17.
35. Nie, J., Patrocinio, A. O. T., Hamid, S., Sieland, F., Sann, J., Xia, S., et al. (2018). New insights
into the plasmonic enhancement for photocatalytic H2 production by Cu-TiO2 upon visible
light illumination. Physical Chemistry Chemical Physics, 20(7), 5264–5273.
36. Atabaev, T. S., Hossain, M. A., Lee, D., Kim, H. K., & Hwang, Y. H. (2016). Pt-coated TiO2
nanorods for photoelectrochemical water splitting applications. Results in Physics [Internet],
6, 373–376. https://doi.org/10.1016/j.rinp.2016.07.002.
37. Leong, K. H., Gan, B. L., Ibrahim, S., & Saravanan, P. (2014). Synthesis of surface plasmon
resonance (SPR) triggered Ag/TiO 2 photocatalyst for degradation of endocrine disturbing
compounds. Applied Surface Science [Internet]., 319(1), 128–135. https://doi.org/10.1016/j.
apsusc.2014.06.153.
38. Tanaka, A., Hashimoto, K., & Kominami, H. (2017). A very simple method for the prepara-
tion of Au/TiO 2 plasmonic photocatalysts working under irradiation of visible light in the
range of 600–700 nm. Chemical Communications [Internet], 53(35), 4759–4762. https://doi.
org/10.1039/c7cc01444a.
39. Abe, R., Takami, H., Murakami, N., & Ohtani, B. (2008). Pristine simple oxides as visible light
driven photocatalysts: Highly efficient decomposition of organic compounds over platinum-­
loaded tungsten oxide. Journal of the American Chemical Society [Internet], 130(25),
7780–7781. Retrieved from https://pubs.acs.org/doi/10.1021/ja800835q.
40. Lin, W., & Frei, H. (2005). Photochemical CO2 splitting by metal-to-metal charge-transfer
excitation in mesoporous ZrCu(l)-MCM-41 silicate sieve. Journal of the American Chemical
Society, 127(6), 1610–1611.
41. Nakamura, R., Okamoto, A., Osawa, H., Irie, H., & Hashimoto, K. (2007). Design of all-­
inorganic molecular-based photocatalysts sensitive to visible light: Ti(IV)−O−Ce(III) bime-
tallic assemblies on mesoporous silica. Journal of the American Chemical Society [Internet],
129(31), 9596–9597. Retrieved from https://pubs.acs.org/doi/10.1021/ja073668n.
42. Irie, H., Miura, S., Nakamura, R., & Hashimoto, K. (2008). A novel visible-light-sensitive
efficient photocatalyst, Cr III-grafted TiO2. Chemistry Letters, 37(3), 252–253.
43. Hush, N. S. (1968). Homogeneous and heterogeneous optical and thermal electron transfer.
Electrochimica Acta, 13(5), 1005–1023.
44. Creutz, C., Brunschwig, B.  S., & Sutin, N. (2005). Interfacial charge-transfer absorption:
Semiclassical treatment. The Journal of Physical Chemistry B [Internet], (109), 10251–10260.
https://doi.org/10.1021/jp050259+.
45. Creutz, C., Brunschwig, B. S., & Sutin, N. (2006, May). Interfacial charge transfer absorp-
tion: Application to metal–molecule assemblies. Chemical Physics [Internet], 324(1), 244–58.
Retrieved from https://linkinghub.elsevier.com/retrieve/pii/S0301010405006610.
46. Irie, H., Miura, S., Kamiya, K., & Hashimoto, K. (2008). Efficient visible light-sensitive
photocatalysts: Grafting Cu(II) ions onto TiO2 and WO3 photocatalysts. Chemical Physics
Letters, 457(1–3), 202–205.
21  Bandgap Engineering as a Potential Tool for Quantum Efficiency Enhancement 563

47. Yu, H., Irie, H., Shimodaira, Y., Hosogi, Y., Kuroda, Y., & Miyauchi, M., et al. (2010, October
7). An efficient visible-light-sensitive Fe(III)-grafted TiO 2 photocatalyst. The Journal of
Physical Chemistry C [Internet], 114(39), 16481–16487. Retrieved from https://pubs.acs.org/
doi/10.1021/jp1071956.
48. Nanaji, K., Siri Kiran Janardhana, R.  K., Rao, T.  N., & Anandan, S. (2019). Energy level
matching for efficient charge transfer in Ag doped-Ag modified TiO2 for enhanced visible
light photocatalytic activity. Journal of Alloys and Compounds, 794, 662–671.
49. Kumar, R., Anandan, S., Hembram, K., & Narasinga, R.  T. (2014). Efficient ZnO-based
visible-light-driven photocatalyst for antibacterial applications. ACS Applied Materials &
Interfaces, 6(15), 13138–13148.
50. Yin, G., Nishikawa, M., Nosaka, Y., Srinivasan, N., Atarashi, D., Sakai, E., et  al. (2015).
Photocatalytic carbon dioxide reduction by copper oxide nanocluster-grafted niobate
nanosheets. ACS Nano, 9(2), 2111–2119.
51. Shoji, S., Yin, G., Nishikawa, M., Atarashi, D., Sakai, E., & Miyauchi, M. (2016). Photocatalytic
reduction of CO2 by CuxO nanocluster loaded SrTiO3 nanorod thin film. Chemical Physics
Letters, 658, 309–314.
52. Raju, K., Navadeepthy, D., Kaliyan, H., Tata, N.  R., & Srinivasan A. (2015, September 1).
Visible-light-induced photocatalytic disinfection of E. coli pathogens with Fe3+-grafted ZnO
nanoparticles. Energy and Environment Focus [Internet], 4(3), 232–238. Retrieved from http://
openurl.ingenta.com/content/xref?genre=article&issn=2326-­3040&volume=4&issue=3&
spage=232.
53. Liu, M., Sunada, K., Hashimoto, K., & Miyauchi, M. (2015). Visible-light sensitive Cu(ii)-
TiO2 with sustained anti-viral activity for efficient indoor environmental remediation. Journal
of Materials Chemistry A, 3(33), 17312–17319.
54. Qiu, X., Miyauchi, M., Sunada, K., Minoshima, M., Liu, M., Lu, Y., et  al. (2012). Hybrid
Cu xO/TiO 2 nanocomposites as risk-reduction materials in indoor environments. ACS Nano,
6(2), 1609–1618.
55. Yu, H., Irie, H., & Hashimoto, K. (2010). Conduction band energy level control of titanium
dioxide: Toward an efficient visible-light-sensitive photocatalyst. Journal of the American
Chemical Society, 132(20), 6898–6899.
56. Qiu, X., Miyauchi, M., Yu, H., Irie, H., & Hashimoto, K. (2010, November 3). Visible-light-­
driven Cu(II)−(Sr 1− y Na y )(Ti 1− x Mo x )O 3 photocatalysts based on conduction band
control and surface ion modification. Journal of the American Chemical Society [Internet],
132(43), 15259–15267. Retrieved from https://pubs.acs.org/doi/10.1021/ja105846n.
57. Anandan, S., & Miyauchi, M. (2011). Ce-doped ZnO (CexZn1-xO) becomes an efficient
visible-­
light-sensitive photocatalyst by co-catalyst (Cu2+) grafting. Physical Chemistry
Chemical Physics, 13(33), 14937–14945.
58. Anandan, S., Ohashi, N., & Miyauchi, M. (2010). ZnO-based visible-light photocatalyst:
Band-gap engineering and multi-electron reduction by co-catalyst. Applied Catalysis B:
Environmental [Internet], 100(3–4), 502–509. https://doi.org/10.1016/j.apcatb.2010.08.029.
59. Kumar, R., Govindarajan, S., Siri Kiran Janardhana, R. K., Rao, T. N., Joshi, S. V., & Anandan,
S. (2016). Facile one-step route for the development of in situ cocatalyst-modified Ti3+ self-­
doped TiO2 for improved visible-light photocatalytic activity. ACS Applied Materials &
Interfaces, 8(41), 27642–27653.
60. Liu, M., Qiu, X., Miyauchi, M., & Hashimoto, K. (2011). Cu(II) oxide amorphous nano-
clusters grafted Ti3+ self-doped TiO2: An efficient visible light photocatalyst. Chemistry of
Materials, 23(23), 5282–5286.
61. Liu, M., Qiu, X., Miyauchi, M., & Hashimoto, K. (2013). Energy-level matching of Fe(III)
ions grafted at surface and doped in bulk for efficient visible-light photocatalysts. Journal of
the American Chemical Society, 135(27), 10064–10072.
62. Liu, M., Qiu, X., Hashimoto, K., & Miyauchi, M. (2014). Cu(ii) nanocluster-grafted,

Nb-doped TiO2 as an efficient visible-light-sensitive photocatalyst based on energy-level
564 R. K. S. K. Janardhana et al.

matching between surface and bulk states. Journal of Materials Chemistry A [Internet], 2(33),
13571–13579. https://doi.org/10.1039/C4TA02211D.
63. Liu, M., Inde, R., Nishikawa, M., Qiu, X., Atarashi, D., Sakai, E., et  al. (2014, July 22).
Enhanced photoactivity with nanocluster-grafted titanium dioxide photocatalysts. ACS Nano
[Internet], 8(7):7229–38. Retrieved from https://pubs.acs.org/doi/10.1021/nn502247x.
64. Inde, R., Liu, M., Atarashi, D., Sakai, E., & Miyauchi, M. (2016). Ti(iv) nanoclusters as a
promoter on semiconductor photocatalysts for the oxidation of organic compounds. Journal of
Materials Chemistry A, 4(5), 1784–1791.
Chapter 22
Nanostructure Material-Based Sensors
for Environmental Applications

Vinutha Srikanth, Mahesh Shastri, M. Sindhu Sree, M. Navya Rani,


Prasanna D. Shivaramu, and Dinesh Rangappa

22.1  Introduction

The sensor for efficient monitoring of the environment has become an emerging
research area over a few decades. These environmental applications include agricul-
tural systems, forests, air quality, weather and storms, volcanoes, and ecological
systems. Efficient environmental sensors with safe and precise measurements for
biological and chemical applications have been realized in recent times. It has been
very obvious these days that a sensor is a feasible option for environmental monitor-
ing like air quality, where the sensor will provide the percentage of gases such as
oxides of carbon, nitrogen, and other toxic gases at ppm levels [1].
Principally, a sensor is a physical device that senses the input from the environ-
ment and converts it into human-readable data. It has become more prominent and
essential in the areas from small consumer products to industrial and military appli-
cations [2, 3]. This sensor is made up of four major components, namely, analyte,
receptor, transducer, and signal processing unit [4]. Here an analyte is a source for
detecting/sensing the defined molecules, and thus the received physical signal is
further converted into an electrical signal by the transducer [5, 6].
Many types of sensors are an actuality in industrial application, and few are
under extensive research. Presently available sensors are built using micro-sized

V. Srikanth
Department of Electrical and Electronics Engineering, KSSEM, Bengaluru, Karnataka, India
M. Shastri · M. Sindhu Sree · P. D. Shivaramu · D. Rangappa (*)
Department of Applied Sciences, Visvesvaraya Technological University,
Center for Postgraduate Studies, Muddenahalli, Chikkaballapur, Karnataka, India
e-mail: dinesh.rangappa@vtu.ac.in
M. Navya Rani
School of Basic and Applied Sciences, Dayanand Sagar University,
Bengaluru, Karnataka, India

© Springer Nature Switzerland AG 2021 565


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_22
566 V. Srikanth et al.

sensing particles and cost-effective. These devices are designed to identify the dis-
tinct unsafe gas molecules in the environment that are venomous when they cross
the edge limit [7]. However, these sensors are suffering from a lack of accuracy and
limited detection level and get contaminates during detection. Thus nanosized par-
ticles have been introduced as a novel sensing material for sensor application. Since
nanomaterials possess a high surface area and surface-to-volume ratio, they are effi-
cient in detecting lower concentration molecules at the PPM (parts per million)/PPB
(parts per billion) level. These advantages of nanomaterials pave the way for a sepa-
rate market for nanomaterial-based toxic gas sensors, which are now gaining atten-
tion among industries and researchers [8].
Nanosensors are broadly classified as optical nanosensors, mechanical nanosen-
sors, or chemical nanosensors [9]. Other than these, the sensors are also classified
based on a range of parameters like the type of nanomaterial, areas of application,
the principle of operation, type of structure, specifications, and means of detection.
One more direction to look at sensors is to consider all the properties of the sensor,
for example, what the sensor is measuring, selectivity range of the sensor, field of
detection, the material used for its fabrication, etc. Optical sensors are guided
through the light rays that travel through the nanomaterial and detect in the range of
wavelength of light [6]. Chemical nanosensors normally detect the type of chemi-
cals present, and the mechanical sensors monitor physical parameters like force,
temperature, etc. Figure 22.1 describes some of the classifications of environmental
sensors based on different applications.
The material that is chosen for a sensor must have the properties, which is
required for the specific application. The various nanomaterials such as metal,

Fig. 22.1  An overview of


the environmental
application of
nanosensors [10]
22  Nanostructure Material-Based Sensors for Environmental Applications 567

semiconductor, carbon, and polymer-based nanoparticle or nanocomposites are


used as a sensor. These materials are synthesized using different physical and chem-
ical methods. The morphology of these nanomaterials depends on the type of syn-
thesis method and various other parameters. Progress in the synthesis and control
over morphology has opened the application of these materials into almost all sec-
tors of life. These sensors are used in environmental monitoring, detecting heavy
ions in water, sensing harmful gases, glucose monitoring, breath analysis, tempera-
ture monitoring, etc.
The evolution of different characteristics of nanomaterials has opened up diverse
applications like semiconductor devices, detection of toxic gases, electrochemical
sensors, biosensors, and many more. Nanomaterials have played a significant role in
the sensor application, which was further developed lab-on-a-chip, in-suit device,
and real-time monitoring system. Here in this chapter few nanostructures such as
carbon, metal, and metal oxide, polymer, and silicon-based nanomaterials are dis-
cussed along with their synthesis, properties, and their application as a sensor in
environmental applications.

22.2  Nanostructured Materials

A nanostructure is defined as its size, where its size ranges from 1 to 100 nm. Hence
they are known as nanostructured materials and possess unique optical, mechanical,
electrical, and magnetic properties [11]. Substantial changes in the properties and
behavior of nanomaterials, when compared to that of bulk materials, have triggered
the interest of many researchers on functionality. The properties of the nanomateri-
als have made them a potential candidate in different fields such as catalysis, bio-
medical, sensors, optoelectronics, energy storage, environmental contamination
detection and remedies, and so on. Among many nanomaterials, carbon, silicon,
metal, and polymer have been significant in use for sensor applications in recent
times. The sensing performance of these materials depends on the microstructures,
size, and bandgap modifications that can be controlled during synthesis.

22.2.1  Carbon-Based Nanomaterials

Carbon-based nanomaterials offer remarkable physical and chemical properties


such as high tensile strength, excellent electrical and thermal properties, stability,
and high corrosion resistance. Hence, carbon nanostructures have been proven as
the best material for sensing toxic gas molecules and other heavy metals. The effi-
ciency and stability of carbon-based sensors have been increased even further when
it is cut down to the nanoscale. Various allotropes of carbon like graphite, diamonds,
and carbon nanostructures like graphene, fullerenes, carbon nanotubes, and gra-
phene quantum dots have given a boost in the area of sensor design and fabrication
568 V. Srikanth et al.

[12]. Due to the simple manufacturing processes, environment compatibility, and


high sensitivity, they are easily combined with other materials to form nanocompos-
ites. These materials are currently being studied and used in different kinds of sen-
sors such as mechanical sensors, pressure sensors, electrochemical sensors, and
biosensors. Graphene-based nanomaterials are extensively used as biosensors for
electrochemical sensing of single- and double-stranded DNA.  Reduced graphene
oxide and its nanocomposites find numerous applications as biosensors, chemical
sensors, and gas sensors. As these nanocomposites are very sensitive, they are used
as a sensing layer in various sensors including plasmonic sensors [13]. Carbon
nanostructures exhibit many advantages compared to the other materials employed
as sensors, especially their environment-friendly behavior with extraordinary physi-
cal and chemical properties.
Nanodiamond is one of the carbon allotropic materials with high biocompatibil-
ity and sensitivity. It has a structure that enables transducing behavior of converting
physical characteristics to optical transitions that can be measured and recognized
in the single-photon range. Nanocomposites based on fullerenes, especially C60, is
used to detect and sense biological molecules like ATP (adenosine triphosphate),
DNA, glucose, licit, and illicit drugs [14]. Graphene when mixed with some poly-
mer is used for strain sensors with increased flexibility and sensitivity. Carbon nano-
tubes and graphene quantum dots are used in biosensing applications [3]. Thus there
are many forms of carbon available for industrial application, and a brief review on
the synthesis of the carbon-based nanomaterials is discussed in the following
sections.

22.2.2  Metal and Metal Oxide-Based Nanomaterials

Metal and metal oxide nanomaterials exhibit unique physical and chemical proper-
ties due to the small-sized, edge surface sites. The metal oxide solid nanoparticles
contain metals as cations and oxide as an anion. Recently, these metal oxides are
emerging as potential materials for gas-sensing and biosensor applications. It has
excellent reactive surface sites that are effective in absorbing targeted molecules.
Few popular metal oxides used for sensor application are titanium oxide (TiO2),
copper (I) oxide (Cu2O), copper(II) oxide (CuO), zinc oxide (ZnO), and many more.
Due to the wide range of bandgaps between the conduction and valence bands,
these nanomaterials are extensively used in almost all types of sensors. Starting
from a small touch sensor to military application and health monitoring, all types of
sensors have been developed [15]. These materials can be prepared by controlling
synthesis conditions to tune the size and morphology that is suitable for the fabrica-
tion of the device. Efforts are being made to use noble metal nanoparticles like gold
and silver as a sensor because of the easy synthesis and fabrication of these metal
nanostructure-based sensors. Besides, they are approachable and easy to handle,
and their composites can be easily procured and processed [16]. These metal nano-
structures can form the building blocks for sensors and other electronic devices
22  Nanostructure Material-Based Sensors for Environmental Applications 569

because of the ability to prepare uniform metal coatings and conductivity detection.
The unique feature of these metal-based nanomaterials as zero-dimensional, one-­
dimensional, and two-dimensional structures has enabled the integration of
nanoscale with biomedical devices [17]. Recent advances in these metal nanoparti-
cles as hybrid structures have resulted in smaller sensing devices for biomedical
applications. These structures with metal oxides are extensively used as sensors for
environmental health monitoring, structural health monitoring, seismic monitoring,
hazardous gas detection, and humidity detection. The textile, food, and agriculture
areas are also adapting these new materials for smart sensing [18]. Recently prog-
ress in zero-dimensional nanocluster preparations has provided potential candidates
for sensor application as they have an exceptionally large surface-to-volume ratio
that can be used for rapid and selective sensing [18].

22.2.3  Silicon-Based Nanomaterials

For decades, silicon has been a promising material for integrated circuit (IC) tech-
niques. Presently, silicon-based nanomaterials are popular due to their excellent
properties and advantage to the conventional IC industry. The silicon-based gas
sensors can cost about $50 each in low volumes, which is very expensive. However,
they have superior optical properties and are usually designed as divergent fluores-
cent sensors and biosensors and, thus, are used in bio-imaging application [19].
Silicon is highly abundant in nature, and its separation techniques are quite easy.
Also, it has a high affinity for oxygen. Due to these properties, silica and silicon-­
based compounds have a special place in high tensile strength and thermal stability
applications. With the advances in silicon technology, silicon-based nanomaterials
are becoming one of the promising members to bridge between nanotechnology and
conventional IC technology. As compared to the pristine polymer, silicon-based
polymer nanocomposites are replacing the new-age materials. Mesoporous silicon
is extensively employed for biosensing and catalytic applications [20]. However, the
synthesis procedure for mesoporous silicon is lengthy and employs tedious pro-
cesses. The composites of silicon with metals, carbon, polymers, gold, and other
noble metals find many applications as sensing agents [21].

22.2.4  Polymer-Based Nanomaterials

Natural polymers are used as biosensors for detecting amines and proteins, whereas
man-made polymers have a great potential for the development of air quality moni-
toring sensors and biosensors [13]. In polymer-based nanocomposites, different
nanostructures are employed as conducting fillers and can absorb the molecules on
the surface of the material. These nanomaterials are mainly used as gas sensors and
humidity-detecting sensors. Conducting polymers have received great attention and
570 V. Srikanth et al.

appreciation in the areas of nanoelectronics, energy storage, biomedical applica-


tions, and sensors. Many conducting polymers are being developed, common among
them are PANI (polyaniline), PPy (polypyrrole), PEDOT (poly 3,4-­ethylene deoxy
thiamine), and polyindole [22]. These are employed for electrochemical sensing
and biosensing applications [23].

22.3  Sensing Mechanism

A sensor is a transducer that transforms any non-electrical input to an electrical


output signal. The transducer is made to pass through different stages before con-
verting into an electrical signal. There are mainly two types of sensors: simple and
complex. Simple sensors give direct output, while the complex sensors are passed
through many processing units before it reaches the final output stage. Understanding
the mechanism of sensing in nanodimensions is important in designing sensors. As
we have seen earlier, nanomaterials are classified as carbon-based, metal-based,
silicon-based, and polymer-based materials. The sensing mechanism in each of
these nanomaterials is different. Also, each type of sensor under each category has
a particular type of sensing mechanism. Figure 22.2 shows the gas-sensing mecha-
nism of NH3 gas sensor [24].
The sensing mechanism in gas sensors involves the adsorption of gas molecules
on the surface of the nanomaterial and subsequently changing the conductance of
the material. Hence, these types of sensors are also called conductometric sen-
sors [25].

Fig. 22.2 A prototype for gas-sensing device (Source credited: Department of


Nanotechnology, VTU)
22  Nanostructure Material-Based Sensors for Environmental Applications 571

There are different operating mechanisms for various nanostructures. The mech-
anism of sensing can be different depending on the type of nanosensor such as
chemical or mechanical. The carbon-based nanostructures are pragmatic in the field
of pressure sensors, electrochemical sensors, optical sensors, and biosensors. They
are not sensitive to a particular type of gas or analyte. Hence, the concept of the
electronic nose was introduced. In the electronic nose, the receptor molecules are
selective to a particular type of odor. These receptor molecules are known as olfac-
tory odor receptors. Each receptor may respond to many numbers of odors. This
prompted the concept of the sensor array. Figure 22.3 shows a schematic mecha-
nism of different types of carbon-based sensors [26–28].
The detection of different types of analytes by carbon-based nanomaterials has
been studied and reviewed by several research groups. Numerous systems are based
on electronic sensors. In addition to these types of electronic and electrochemical
sensors, optical sensors are gaining importance. All types of analytes are detected
using optical sensors based on CNT and graphene-nanostructured sensing materi-
als. Among the entire range of carbon-based sensors, the DNA-modified sensors are
the most prominent one. The recent advancement in CNT-based sensors is the car-
bon nanohorns, which are applicable for detecting impurities in metals and glucose
monitoring.
Biological sensors work on a different mechanism. Jeon et al. have mentioned
that the adsorbed molecules are further converted as potential carriers, and the per-
mittivity, as well as refractive index, is changed. Figure  22.4 shows the sensing
mechanism of the biological sensor.
Voltammetric sensors work on the principle of change in current for varying volt-
age. Another type of sensor is potentiometric sensors, wherein the sensing mode is
based on the chemical potential of the system which is altered by the presence of an
analyte when no current is flowing. Amperometric sensors work on a similar mech-
anism to voltammetric sensors, which have a principle of measuring current that is
generated by the redox reactions of the analyte near the electrodes. Apart from
these, several sensing mechanisms are depending on the type of sensors like the
FET-based sensors, transistor sensors, capacitive sensors, etc. All biosensors work

(a) Electronic Nose sensor


(b) FET based Sensor
laser
attached OR graphene
Cr/Au electrode
NT

SiO2
Si

Fig. 22.3  Schematic of carbon-based nanosensors: (a) electronic nose sensor and (b) FET-­
based sensor
572 V. Srikanth et al.

Fig. 22.4  Schematic of sensing mechanism of biosensor

on a different mechanism of placing the samples in a liquid medium [29]. The ana-
lyte in a biosensor can be DNA strand, blood, urea, uric acid, any type of amine,
peptide, etc.

22.4  Synthesis of Nanostructures for Sensor Materials

22.4.1  Carbon-Based Nanostructures

Carbon nanotubes (CNTs) are synthesized by various methods that are used keeping
the desired properties in mind that are required for specific applications.

Chemical Vapor Deposition (CVD)

Multiwalled carbon nanotubes (MWCNTs) can be synthesized by chemical vapor


deposition (CVD) with Fe1−xCox as a catalyst supported by group II elements as
shown in Fig. 22.5a. In this method, a suitable amount of catalyst with 5 weight
percent of group II elements M-CO3 (M = Ca, Mg, Ba, Sr) is used to perform the
reaction for 30  min at different temperatures. This shows that various processes
have been deployed for the growth of CNTs, in particular for most efficient and low-­
temperature nanotube growth [30]. CaCO3 was used as a catalyst for increasing the
CNT yield. The highest yield of CNT at 33 mol% catalyst was achieved by J W
Seo et al.
The plasma-enhanced chemical vapor deposition (PECVD) can also be used to
synthesize CNT [31]. Ebbesen et  al. have developed a large-scale production of
CNT by microwave plasma-enhanced chemical vapor deposition (MPECVD) and
22  Nanostructure Material-Based Sensors for Environmental Applications 573

Fig. 22.5 (a) Schematic of CCVD. (b) TEM images of nanotubes grown by CVD method [30]

CVD [32]. The experiments are usually carried out at atmospheric pressure in the
flow path. There are two different types of furnaces that are being used in
PECVD. One is a horizontal type and the other one is a vertical type. The use of a
horizontal furnace is presiding, where the catalyst is placed into quartz or ceramic
boat and placed in a long quartz tube. The reaction mixture consists of inert gas,
hydrocarbon gas, and catalyst that are loaded in the tube. The mixture containing a
source of an inert gas such as Ar or N2 and hydrocarbon in the form of gas is purged
over the catalyst bed at temperatures from 500 to 1100 °C. The vertical-type furnace
configuration is commonly used for large-scale production of carbon nanotubes/
carbon fibers and the source; the catalyst is introduced at the top of the furnace at
industry levels. The resulting filament accumulates in flight and then on the bottom
of the chamber. The particles of the ultrafine metal catalyst are introduced directly
into the reactor or formed in situ using a precursor such as a metallocene [33, 34].
Both single-walled nanotube and multiwalled carbon nanotube (SWCNT and
MWCNT) nanostructures can be synthesized using different techniques, where one
can control the size of the nanotube by changing the active compounds on the sur-
face of the catalyst. The reaction time decides the length of the tube; even long tubes
of up to 60  mm can be developed [35]. The catalytic decomposition of carbon
source with helium and hydrogen as carrier gases at 1150  °C can lead to the
SWCNTs with a diameter of 3.23 nm [36]. Yah et al. have synthesized the CNT at
900–950 °C by using ferrocene as a catalyst and compared it with 800–850 °C CNT,
which were further preferred to industrial production [37].

Laser Ablation

Another process for manufacturing carbon nanotubes is the laser ablation method
[35]. In this method, carbon is made to evaporate from the surface of graphite with
high-density argon using graphite-focused pulse laser. A graphite substrate is placed
in the center of a 60-mm quartz tube rising in a tubular furnace shown in Fig. 22.6.
After the sealed tube has been removed, the furnace temperature is increased to
1200 °C. The Gaussian output beam lens with a focal length of 75 or 200 cm is used
to create a 3-mm or 6-mm beam spot on each target. The flow of argon carrier gas
574 V. Srikanth et al.

Fig. 22.6  Schematic representation of the oven laser vaporization apparatus used at Rice
University (Houston, Texas, USA) [39]

transports most of the carbon obtained from the process of laser evaporation and
high-temperature condensation setup present in the furnace section and precipitates
the carbon in the form of soot on a water-cooled conical copper rod [38].
Recently, Ismail et al. synthesized carbon nanoparticles (CNPs) and MWCNTs
by using pulsed laser ablation of a graphite target in water without using cata-
lyst [38].

Arc Discharge

Sumio Iijima has demonstrated the arc-discharge evaporation method for the first
time that was used for fullerene synthesis [40].
The latest version of the arc-discharge method is shown in Fig. 22.7. Graphite
rods act as electrodes: cathode and anode. In this process, a high current is applied
around 50–100 A through two graphite electrodes in a closed chamber. High tem-
perature (>1700  °C) is ignited through plasma of inert gas at low pressure
(50–700 mbar) into the chamber [42, 43]. Two graphite electrodes with an average
diameter of about 6–12 mm are faced at intervals of 1–4 mm. Carbon atoms evapo-
rated due to high temperature and high pressure form a small carbon cluster at
anode graphite rod (C3/C2) [42]. As a result, this cluster cathode electrode precur-
sor or metal surface catalyst rearranges the surface into microtubules such as CNT
rescue. However, the desire to form CNT (SWCNT or MWCNT) mainly depends
on inert gas, temperature, current, and catalyst exhaust chamber. The SWCNT is
relatively pure in the arc-discharge method. In the arc-discharge method, frequently
used catalysts are Ni, Co, Fe, and Pt, and very often they are used as a catalyst (i.e.,
mixed catalysts) for higher SWCNT yield.
22  Nanostructure Material-Based Sensors for Environmental Applications 575

Fig. 22.7  Schematic representation of methods used for carbon nanotube synthesis. (a) Arc dis-
charge. (b) Chemical vapor deposition. (c) Laser ablation. (d) Hydrocarbon flames [41]

22.4.2  Metal-Based Nanoparticles

In sensors, the main interest is the design and development of nontoxic multifunc-
tional nanomaterials to meet future application needs. There have been more reports
on semiconductor nanomaterials gaining importance over the past few years in cata-
lysts, medical, food industry, health monitoring, diagnostics, energy storage, and
environmental fields [44]. Metal oxide-based sensors can be easily developed with
a variety of simple manufacturing methods, and they offer high sensitivity at a low
cost [45]. Due to high adsorption capacity, large surface area, stability, and unique
electrochemical activity, metal oxide nanomaterials are very important for design-
ing a platform for electrochemical sensors [46].
Various types of nanomaterials, including single or hybrid/combination nano-
structures, are being designed to have outstanding features different from ordinary
bulk materials. The metal oxide-based nanostructures can be synthesized by differ-
ent methods. The most commonly used method is a solution method where metal
salts dissolved in solution are precipitated to metal hydroxide followed by dehydra-
tion reaction with the increase in the reaction temperature and time. The solution
method offers many advantages in controlling the size and morphology of metal
oxide nanoparticles. For example, nanofiber-, nanorod-, nanosphere-, and nanotube-­
like morphology can be easily obtained by the chemical reduction method. Among
the different nanostructures, the 1D nanostructure provides a detecting system for
the electrochemical detection of harmful gases. Some of the semiconductors such as
ZnO, TiO2, SnO2, In2O3, NiO, V2O5, and WO3 play a key role in detection and envi-
ronmental monitoring of explosive/toxic gases [47].
576 V. Srikanth et al.

Recently, Qingji Wang et al. successfully synthesized In-Sn composite by one-­


pot hydrothermal method. This metal nanocomposite showed an excellent sensing
response to ethanol gas. The improvement of the gas-sensing reaction is explained
due to the smaller-sized nanoparticles, heterojunction formed between indium oxide
and tin oxide, and the redox ability of the In-Sn oxide-based complex. The TEM
images of indium oxide and tin oxide nanoparticles are shown in Fig. 22.8 [48].
Lee et al. reported αFe2O3 nanoparticle-decorated ZnO nanowire synthesized by
vapor-liquid-solid (VLS) process and sol-gel process was employed as CO gas sen-
sor. This was achieved by preparing the nanowires via the VLS process. The catalyst
in the form of liquid droplets on the substrate was used by purging N2/O2 gas. Sn
with ITO glass and Sn powder acts as a catalyst for ZnO NW’s growth. The tem-
perature was increased up to 900 °C while purging 100 ppm of N2 gas as shown in
Fig.  22.9 [27]. The Fe2O3 Nanoparticle was synthesized by sol-gel technique,

Fig. 22.8  TEM images of In-Sn oxide composite [48]


22  Nanostructure Material-Based Sensors for Environmental Applications 577

Fig. 22.9 (a) Schematic of the furnace and experimental equipment. (b) Zn-Sn powder on ITO via
VLS method

Fig. 22.10  Preparation of


SnO2/ZnO hierarchical
nanostructures [49]

deposited on the ZnO substrate by spin coating, and heated up to 400 °C under N2
atmosphere for 1 h [27].
Muhammad Tahir Zahoor et  al. developed rod-like WO3 nanostructures via
hydrothermal synthesis at 180 °C for 48 h. The morphology of the WO3 nanostruc-
tures was examined by TEM, and it shows rod-like morphology with hexagonal
crystalline structure as confirmed by XRD [44].
Khoang et al. designed hierarchical SnO2/ZnO nanostructures that were prepared
by thermal evaporation and hydrothermal process as high-performance volatile eth-
anol sensors. The SnO2 was synthesized in the horizontal-type furnace at 800 °C
under O2 atm as an Au catalyst. Then the prepared SnO2 NWs were coated on the
ZnO by spray coating under heat treatment at 300  °C.  Steps of preparation are
shown in Fig. 22.10 [49].
578 V. Srikanth et al.

22.4.3  Polymer-Based Nanoparticles

In recent studies, conductive polymers are being used as an alternative to metal


oxide for sensing applications. Presently, the conjugated polymer has attracted great
attention for gas detection applications. Conjugate polymer-based sensors have
great advantages over solid-state sensors. High selectivity and sensitivity at room
temperature require a low device voltage, large-scale manufacturing, low cost, and
ease of application [50]. Bentonite nanohybrid-modified polyaniline (PANI) nano-
fibers have been used in gas sensors in the detection of harmful gas molecules such
as benzene, acetone, ethanol, and toluene. The Ru and Ag electrodes were used as a
substrate to coat bentonite nanohybrid by the dip-coating method and employed to
investigate the toxic gases like toluene, benzene, ethanol, and acetone [51]. Ji Young
Lee et al. have developed a simple and selective method for electrochemical detec-
tion of hydrazine using poly(dopamine) (pDA)-modified In-Sn on ITO electrodes.
The hydrogen detection was carried out by cyclic voltammetry to determine the
concentration of H2 on the pDA in the range of 100 μM–10 mM at 0.3 V. It does not
affect the ionic species when it is used to analyze tap water, and it shows very good
recovery [52]. Navale et al. developed CSA-doped PPy/⍺-Fe2O3 hybrid nanocom-
posites by solid-state synthesis method. The nanocomposite was deposited on a
glass substrate under the influence of 100 ppm of various oxidation and reducing
gases such as NO2, Cl2, H2S, NH3, C2H5OH, and CH3OH using room temperature
two-probe resistance measurement setup [53]. It shows a very fast response time,
good stability, and better performance than PPy at room temperatures.
Wang et al. fabricated an electrochemical sensor for the detection of Hg2+ with
ultrasensitivity. Here, AuNP-BSA-rGO hybrid nanocomposite was used as a sub-­
surface material for immobilization of triple-stranded DNA as shown in Fig. 22.11.

22.5  Applications

22.5.1  Toxic Gas Sensor

An increased developmental activity, populations, production of energy, and indus-


tries cause the release of huge amounts of toxic gases as by-products like carbon
monoxide, hydrogen oxide, hydrogen chloride, benzene, sulfides, and many more.
These toxic gases have become significantly dangerous to our health and the envi-
ronment. Therefore, it needs an hour for the detection and reduction of toxic gas
percentages in the environment.
Gas sensors have a great influence in many areas such as environmental monitor-
ing, domestic safety, public security, automotive applications, and air conditioning
in airplanes, spacecraft, houses, and sensor networks. The huge application range of
these sensors have demanded the need for developing cheap, small, low-power-­
consuming, and reliable solid-state gas sensors. Hence, over the years, a huge
22  Nanostructure Material-Based Sensors for Environmental Applications 579

Fig. 22.11  Preparation of the electrochemical probe for Hg2+ detection [54]

research activity has been conducted worldwide to overcome drawbacks in the


metal oxide-based sensors, to sum up in improving the well-known “3S” sensitivity,
selectivity, and stability [55]. Most studies have focused on detecting CO, CO2, SO2,
O3, O2, H2, gases, and some of the organic vapors such as methanol, ethanol, isopro-
panol, and benzene. Mostly, these studies have been carried out in a simple matrix,
containing inert atmosphere (Ar, N2), or in synthetic air. Therefore, the required
selectivity for certain gases in the presence of complex samples and matrices is yet
to be evaluated [56]. The sensor properties of semiconductor metal oxides for thin
films or thick films other than SnO2 such as TiO2, ZnO, WO3, In2O3, and Fe2O3 are
also being studied. Noble metals such as Pt, Pd, Ag, and Au improve material selec-
tivity and stability [57].
(a) Metal Oxide-Based Nanomaterials: Nanostructured metal oxides are one of
the main types of materials used in gas-sensor device fabrication. Among these
commonly used metal oxides are semiconductors. One of the electrical proper-
ties of these metal oxides is change in temperature. Nanostructured materials
possess a high surface-area-to-volume ratio which in turn enhances the number
of active sites further aiding in improved sensor performance [57]. Jeong Seok
Lee et al. have reported the synthesis of a hierarchical structure of ⍺-Fe3O4-ZnO
nanowire by VLS and sol-gel method for gas sensor applications. Here they
have coated this sensor material on the surface of In-Sn oxide as conductive
substrate. Further, the gas-sensing property was analyzed using CO, NH3, and
NO2 at varying temperatures and varying concentrations for which the response
of CO had relatively improved [27].
Recently, Qingji Wang et al. reported the synthesis of environmentally friendly
In-Sn oxide composite using a one-pot hydrothermal method. They have
580 V. Srikanth et al.

investigated the gas-sensitive properties of ethanol along with the mechanism. The
reaction of the In-Sn oxide complex per 100 ppm of ethanol was 59.6, which was
7.9 times higher than tin oxide at 200 °C shown in Fig. 22.12 [58].
Manish Kumar Verma et  al. reported the sensor response properties of SnO2-­
CuO multilayer to H2S at varying volume concentrations (2–20 ppm). It has been
found that the SnO2-CuO multilayer structure with the optimum 3% by volume
CuO exhibits a high response of 2.7 × 104 within a response time of 2 s (t90), oper-
ated at 140 °C as shown in Fig. 22.13 [59].
Mohamed A. Basyooni et al. reported the synthesis of Na-doped ZnO membrane
having predominant orientation along the C axis using the sol-gel method. The ZnO
bond length increased from 1.980 to 1.986 Å at 2.5% Na. The wrinkle network-like
nanostructure was obtained and observed a very high sensitivity of 81.9% at 50
sccm, for CO2 gas. Further, both reaction time and recovery time have been increased
from 179 to 240  s and from 122 to 472 [60]. An experimental setup of the gas-­
sensing system is shown in Fig. 22.14.
(b) Carbon-Based Nanostructure: Jin Wu and co-workers synthesized 3D hierar-
chical graphene oxide and reduced graphene oxide structures by using one-step
spark plasma sintering (SPS) for detection of NO2 gas. Using spark plasma
sintering technique, most of the oxygen-containing functional groups can be
removed by heat reduction process and high vacuum spark. With the property of
superhydrophobicity with rGO, improved the surface area with large defects
that promotes gas adsorption. Investigated the application of the NO2 sensor and
exhibited high sensitivity at 25.5 ppm with a low detection limit to 9.1 ppb [61].
Keshtkar et al. developed a novel sensing material of nanohybrid SnO2 quantum
dots-fullerene (SnO2 QDs-C60) using a hydrothermal process. The gas-sensing
property was analyzed for air and H2S with 70 ppm at different temperatures.
The sensor shows a higher and faster response for H2S gas [62]. Algadri et al.
have fabricated an H2 gas sensor by using the dielectrophoresis method. Here
CNT has been used as a sensing material. The CNT was synthesized using the

Fig. 22.12  Based on the sample at 200  °C shows responses of the sensor: (a) three cycling
response transient curves and (b) the long-term stability
22  Nanostructure Material-Based Sensors for Environmental Applications 581

Fig. 22.13  Fabricated sensor: (a) SnO2 substrate and (b) composite multilayered structure. (c)
The graph shows the variation of the response of sensor. Inset as a function of H2S gas varia-
tions [59]

Fig. 22.14  The experimental section of the gas-sensing system [60]

mixture of graphite and ferrocene in the microwave. The sensor is exposed to a


20–1000 ppm mixture of H2-N2 sensor which shows high sensitivity up to 240%
upon exposure to 1000 ppm of H2 at room temperature (RT) [63].
Zhao and co-workers investigate the diatomic gas detection of CO, O2, and NO
by C20 molecular junctions. They have theoretically calculated the adsorption energy
and transport properties of C20 using nonequilibrium Green’s function (NEGF) for-
malism in combination with density functional theory (DFT). C20 is limited to CO
perception, but the results showed that NO and O2 can be selectively detected [64].
Similar work was carried out by Rahimi et al. Sc-doped C20 (ScC19) nanocage is
theoretically using density functional theory (DFT). The result shows that for rare
gas adsorption, the Sc-doped C20 molecule has significantly changed its electronic
properties [65].
582 V. Srikanth et al.

22.5.2  Humidity Sensor

Humidity determines the amount of water in the vapor-phase measure in a gas sam-
ple. The development of sensors at low cost for accurate humidity measurements is
essential for application in the agriculture, chemical, food production and storage,
semiconductor, textile, and pharmaceutical industry [66]. There are several tech-
niques for designing humidity sensors: resistive, mechanical, capacitive, optical,
and acoustic techniques for transduction [66].
Dubourg et al. have reported TiO2 nanoparticles as a sensing material for humid-
ity sensors, where the sensor was designed on the PET substrate by screen printing
method as shown in Fig. 22.15. The fabricated sensor was operated at room tem-
perature, i.e., 25 °C. Linear response-sensitive layers made in the range of relative
humidity from 5 to 70% show a great potential for environmental monitoring and
humidity measurement [67]. In popularity, the humidity sensors are fabricated by
three-dimensional graphene foam (3DGF) nanostructures. Leng et al. developed a
modified graphene oxide (MGO)/nafion hybrid sensor for measuring the humidity.
A long-term stability test shows that nafion in MGO can improve the stability of
MGO/nafion sensors, while too much of it will decrease its linearity [68]. Quartz
microbalance (QCM) is used as a humidity sensor in various types of sensors due to
its better resolution, cost-effectiveness, and digital output frequency. The QCM
moisture sensors typically apply electrode-sensitive films [69, 70]. Hens et al. pre-
pared the sensor based on nanodiamond particles to disperse CNT and hybrid nano-­
carbon-­based composites [71].
Similar work was carried out by Li et al. who studied MWCNT/GO composite
fabricated by drop-casting method for humidity detection. The prepared MWCNT/
GO composite was coated on the QCM. The addition of MWCNT increases the
distance between the GO sheets of the composite and may be useful for the diffu-
sion of water molecules. This MWCNTs/GO composite shows a great potential for
highly sensitive humidity detection [72]. In another study, Ding et al. had reported
fullerene C60/GO-based nanocomposite and fabricated QCM-based humidity sen-
sors by using these hydrophilic functional groups. This enhanced the sensitivity of
the sensor. The deposition of the nanocomposite was carried out by drop-casting on

Fig. 22.15  Shows the process and design of resistive-type chemical sensor [67]
22  Nanostructure Material-Based Sensors for Environmental Applications 583

the QCM electrode. However, interacting with water molecules can increase the
viscosity of GO molecules and damage the sensor. After evaluating different sen-
sors, the C60/GO QCM humidity sensor was found better due to its good response,
recovery, and stability [69]. Saha et al. reported moisture sensors by gel tape-casting
thick films of fullerene (C60)-modified nanoporous ɣ-alumina (Al2O3). It shows the
excellent capability and high sensitivity for moisture sensors in the range of
1–25 ppm for various industrial applications [73].

22.5.3  Biosensors

Biosensor plays an important role in environmental monitoring, food packaging,


and biosafety. They are capable of detecting genetic markers of genetic abnormali-
ties, pathogens, viruses, toxins, and diseases [74]. Among all these electrochemical
biosensors take great attention because of their sensitive, selective, cost-effective,
and fast response time. Few examples of electrodes used in electrochemical biosen-
sors are gold(Au) nanoparticles, carbon (C), electrically conducting polymer (CP),
and carbon nanotubes (CNTs) [75]. Sharif Ahmad and co-workers studied the dis-
persion of polymer hydrophobic coating of nano-conducting polymer (CP).
Conducting polymers (CP) such as polyaniline, polypyrrole, polythiophene (PTh),
and its derivatives are considered as good materials for biosensors [76].
Hongxiu Dai et  al. developed an electrochemical sensor using phytic acid-­
functionalized polypyrrole (PPy)/graphene oxide (GO) and modified electrodes for
the detection of heavy metal ions. In this study, Cd (II) and Pb (II) were measured
using phytic acid-modifying electrodes based on PPy/GO nanocomposites. The
modified PA/PPy/GO electrode also showed high reproducibility and stability in
acetic acid-sodium acetate buffer. With the advent of more functional materials for
analytes, the proposed strategy seems to provide a potential platform for analyzing
and processing trace amounts of heavy metals in the environment [77].
Wang et al. developed an ultrasensitive electrochemical sensor for the detection
of Hg+2. They have synthesized Au-BSA-rGO composite for immobilization of
triple-­helix DNA. In the presence of Hg+2, a stable T-Hg+2-T complex was generated
which destroyed the triple-helix DNA, leading to the combination of the free sDNA
and the cDNA on AuNP [54]. Heng et al. developed a Hg2+ detection using hybrid-
ized DNA as a capping scaffold based on the fluorescent Ag nanocomposite. This
composite provides the detection of Hg2+ (aqueous) using duplex DNA with Ag
nanocomposite as probes [78]. Farias et al. developed an electrochemical sensor for
the determination of anticancer drug flutamide in pharmaceutical and artificial urine
samples. Here cyclic voltammetry technique was used for the analysis of the cata-
lytic response of flutamide and was prepared by using a glassy carbon electrode for
varying concentrations [79]. Jeong et  al. synthesized 3D nanocomposite N-GR-­
CNTs/AuNP biosensors for non-enzymatic glucose detection. It shows good behav-
ior for glucose detection with a range of 2 μM to 19.6 mM and good sensitivity of
0.9824 μA.m/M.cm−2 [80]. Kannegulla et al. fabricated nanosensor by quenching
584 V. Srikanth et al.

property of C60 fullerenes with quantum dots (QD) using charge transfer process.
This property allows the creation of a molecular beacon (MB) sensor consisting of
a QD reporter and multiple C60 label probes. The results indicate the high selectiv-
ity of the MB at molecular nanoparticle (MNP) probe has a great potential in low
concentration DNA detection without polymerase chain reaction (PCR) [81].

22.6  Summary and Outlook

A large number of researches and studies have shown that sensors play an important
role in the industrial, agricultural, food, commercial, military, biological, and many
other sectors of human interventions. Through this, it can be seen that carbon-based
nanomaterials and metal-based nanomaterials are widely used in the development
of sensors. Though there are a number of sensors in the market with good sensitivity
and selectivity, refractive-index, functionalized nanomaterials cannot be bulk manu-
factured. There is no unique method that can be adopted and carried to manufacture
nanosensors. A great amount of effort has been made to overcome the defects and
shortcoming of the parameters related to sensors. Detecting target species in a com-
plex environment is still a challenge. Another concern is that the nanomaterials
degrade over a while. Hence many sustainable materials have to be studied and
implemented for building sensors and sensor subsystems. We have summarized four
types of nanomaterials keeping in view the field of application. Gas sensors, envi-
ronmental monitoring, biosensors for glucose, and DNA monitoring sensors are few
application areas we have focused on. Future research must focus on the construc-
tion of next-generation sensor devices. It is anticipated that high-performance sen-
sors with high stability and smart performance will be developed and
manufactured.

References

1. Tissa, A.P.J., Illangasekare, H., & Han, Q. (2018). Environmental underground sensing and
monitoring.
2. Vogi F., Dable B., Cramer J., Books K. (2009), Recent advances in Chemometrics for Smart
sensors, Analyst, 129, 492–502.
3. Lim, T.  C., & Ramakrishna, S. (2006). A conceptual review of nanosensors. Zeitschrift für
Naturforschung A—A Journal of Physical Sciences, 61, 402–412. https://doi.org/10.1515/
zna-­2006-­7-­815.
4. Fraden, J., & King, J. G. (1998). Handbook of modern sensors: Physics, designs, and applica-
tions (2nd ed.).
5. Abdel-Karim, R., Reda, Y., & Abdel-Fattah, A. (2020). Review—Nanostructured materials-­
based nanosensors. Journal of the Electrochemical Society, 167, 037554. https://doi.
org/10.1149/1945-­7111/ab67aa.
6. Bogue, R., & Bogue, R. (2008). Nanosensors: A review of recent progress. Sensor Review, 28,
12–17. https://doi.org/10.1108/02602280810849965.
22  Nanostructure Material-Based Sensors for Environmental Applications 585

7. Lupan, O., Pauporté, T., & Viana, B. (2010). Low-voltage UV-electroluminescence from ZnO-­
nanowire array/p-GaN light-emitting diodes. Advanced Materials, 22, 3298–3302. https://doi.
org/10.1002/adma.201000611.
8. Šutka, A., & Gross, K. A. (2016). Spinel ferrite oxide semiconductor gas sensors. Sensors and
Actuators B: Chemical, 222, 95–105. https://doi.org/10.1016/j.snb.2015.08.027.
9. Abdel-Karim, R., Reda, Y., & Abdel-Fattah, A. (2020). Review—Nanostructured materials-­
based nanosensors. Journal of the Electrochemical Society, 167, 37554. https://doi.
org/10.1149/1945-­7111/ab67aa.
10. Springer. (2020). Nanosensor technologies for environmental monitoring.
Retrieved from https://link.springer.com/bookseries/15921%0Ahttp://link.springer.
com/10.1007/978-­3-­030-­45116-­5.
11. Expanding the Vision of Sensor Materials. (1995). https://doi.org/10.17226/4782.
12. Gomez, I. J., Arnaiz, B., Cacioppo, M., Arcudi, F., & Prato, M. (2018). Nitrogen-doped Carbon
Nanodots for bioimaging and delivery of paclitaxel. Journal of Materials Chemistry B, 6.
https://doi.org/10.1039/x0xx00000x.
13. Semwal, V., & Gupta, B. D. (2019). Highly sensitive surface plasmon resonance based fiber
optic pH sensor utilizing rGO-Pani nanocomposite prepared by in situ method. Sensors and
Actuators B: Chemical, 283, 632–642. https://doi.org/10.1016/j.snb.2018.12.070.
14. Bezzon, V.  D. N., Montanheiro, T.  L. A., De Menezes, B.  R. C., Ribas, R.  G., Righetti,
V. A. N., Rodrigues, K. F., & Thim, G. P. (2019). Carbon nanostructure-based sensors: A brief
review on recent advances. Advances in Materials Science and Engineering, 2019. https://doi.
org/10.1155/2019/4293073.
15. Schneider, K., & Maziarz, W. (2018). V2O5 thin films as nitrogen dioxide sensors †. Sensors,
18. https://doi.org/10.3390/s18124177.
16. Harold H. Szu, F. Jack Agee (2009), Independent Component Analyses, Wavelets, Neural
Networks, Biosystems, and Nanoengineering VII, Proceedings of SPIE, volume 7343, 734301
17. Tripathy, N., & Kim, D.-H. (2018). Metal oxide modified ZnO nanomaterials for biosensor
applications. Nano Convergence, 5, 27. https://doi.org/10.1186/s40580-­018-­0159-­9.
18. Zhang, B., & Gao, P. X. (2019). Metal oxide nanoarrays for chemical sensing: A review of
fabrication methods, sensing modes, and their inter-correlations. Frontiers in Materials, 6.
https://doi.org/10.3389/fmats.2019.00055.
19. Elosua, C., Arregui, F.  J., Del Villar, I., Ruiz-Zamarreño, C., Corres, J.  M., Bariain, C.,
Goicoechea, J., Hernaez, M., Rivero, P. J., Socorro, A. B., Urrutia, A., Sanchez, P., Zubiate,
P., Lopez-Torres, D., De Acha, N., Ascorbe, J., Ozcariz, A., & Matias, I. R. (2017). Micro and
nanostructured materials for the development of optical fibre sensors. Sensors, 17. https://doi.
org/10.3390/s17102312.
20. Ahmed, W., Subramani, K., & Elhissi, A. (2019). Introduction to nanotechnology.

Nanobiomaterials in Clinical Dentistry, 3–18.
21. Tieu, T., Alba, M., Elnathan, R., Cifuentes-Rius, A., & Voelcker, N. H. (2019). Advances in
porous silicon-based nanomaterials for diagnostic and therapeutic applications. Advances in
Therapy, 2, 1800095. https://doi.org/10.1002/adtp.201800095.
22. Di Zhang, H., Tang, C. C., Long, Y. Z., Zhang, J. C., Huang, R., Li, J. J., & Gu, C. Z. (2014).
High-sensitivity gas sensors based on arranged polyaniline/PMMA composite fibers. Sensors
and Actuators A: Physical, 219, 123–127. https://doi.org/10.1016/j.sna.2014.09.005.
23. Wang, G., Morrin, A., Li, M., Liu, N., & Luo, X. (2018). Nanomaterial-doped conducting
polymers for electrochemical sensors and biosensors. Journal of Materials Chemistry B, 6,
4173–4190. https://doi.org/10.1039/c8tb00817e.
24. Chauhan, P. (2016). Nanomaterials for sensing applications. Journal of Nanomedicine

Research, 3, 1–8. https://doi.org/10.15406/jnmr.2016.03.00067.
25. Rout, C. S., Hegde, M., Govindaraj, A., & Rao, C. N. R. (2007). Ammonia sensors based on metal
oxide nanostructures. Nanotechnology, 18. https://doi.org/10.1088/0957-­4484/18/20/205504.
586 V. Srikanth et al.

26. Baptista, F. R., Belhout, S. A., Giordani, S., & Quinn, S. J. (2015). Recent developments in car-
bon nanomaterial sensors. Chemical Society Reviews, 44, 4433–4453. https://doi.org/10.1039/
c4cs00379a.
27. Lee, J., Lee, S.-H., Bak, S.-Y., Kim, Y., Woo, K., Lee, S., Lim, Y., & Yi, M. (2019). Improved
sensitivity of α-Fe2O3 nanoparticle-decorated ZnO Nanowire Gas Sensor for CO. Sensors
(Basel), 19. https://doi.org/10.3390/s19081903.
28. Heddle, J.  G. (2008). Protein cages, rings and tubes: Useful components of future nanode-
vices? Nanotechnology, Science and Applications, 1, 67–78.
29. Riu, J., Maroto, A., & Rius, F. X. (2006). Nanosensors in environmental analysis. Talanta, 69,
288–301. https://doi.org/10.1016/j.talanta.2005.09.045.
30. Seo, J.  W., Magrez, A., Milas, M., Lee, K., Lukovac, V., & Forró, L. (2007). Catalytically
grown carbon nanotubes: From synthesis to toxicity. Journal of Physics D: Applied Physics,
40, R109–R120. https://doi.org/10.1088/0022-­3727/40/6/r01.
31. Li, Y., Mann, D., Rolandi, M., Kim, W., Ural, A., Hung, S., Javey, A., Cao, J., Wang, D.,
Yenilmez, E., Wang, Q., Gibbons, J. F., Nishi, Y., & Dai, H. (2004). Preferential growth of
semiconducting single-walled carbon nanotubes by a plasma enhanced CVD method. Nano
Letters, 4, 317–321. https://doi.org/10.1021/nl035097c.
32. Anon. (1992). No Title. Nat. Publ. Gr.
33. Rahman, G., Najaf, Z., Mehmood, A., Bilal, S., ul Haq Ali Shah, A., Mian, S.  A., & Ali,
G. (2019). An overview of the recent progress in the synthesis and applications of carbon
nanotubes. C, 5. https://doi.org/10.3390/c5010003.
34. Kolmakov, A., & Moskovits, M. (2004). Chemical sensing and catalysis by one-dimensional
metal-oxide nanostructures. Annual Review of Materials Research, 34, 151–180. https://doi.
org/10.1146/annurev.matsci.34.040203.112141.
35. Journet, C., & Bernier, P. (1998). Production of carbon nanotubes. Applied Physics A:

Materials Science & Processing, 67, 1–9. https://doi.org/10.1007/s003390050731.
36. Kasperski, A., Weibel, A., Datas, L., De Grave, E., Peigney, A., & Laurent, C. (2015). Large-­
diameter single-wall carbon nanotubes formed alongside small-diameter double-walled car-
bon nanotubes. Journal of Physical Chemistry C, 119, 1524–1535. https://doi.org/10.1021/
jp509080e.
37. Yah, C. S., Simate, G. S., Moothi, K., Maphutha, K. S., & Iyuke, S. E. (2011). Synthesis of
large carbon nanotubes from ferrocene: The chemical vapour deposition technique. Trends in
Applied Sciences Research, 6, 1270–1279.
38. Ismail, R. A., Mohsin, M. H., Ali, A. K., Hassoon, K. I., & Erten-Ela, S. (2020). Preparation
and characterization of carbon nanotubes by pulsed laser ablation in water for optoelectronic
application. Physica E: Low-Dimensional Systems and Nanostructures, 119, 113997. https://
doi.org/10.1016/j.physe.2020.113997.
39. Guo, T., Nikolaev, P., Rinzler, A. G., Tombnek, D., Colbert, D. T., & Smalley, R. E. (1995).
Self-assembly of tubular fullerenes. 10694–10697.
40. Ando, Y., & Iijima, S. (1993). Preparation of carbon nanotubes by arc-discharge evaporation.
Japanese Journal of Applied Physics, 32. https://doi.org/10.1143/JJAP.32.L107.
41. Gore, J. P., & Sane, A. (2011). Flame synthesis of carbon nanotubes. In S. Yellampalli (Ed.),
Carbon nanotubes. Rijeka: IntechOpen. https://doi.org/10.5772/21012.
42. Ahmad Aqel, Kholoud M. M. Abou El-Nour, Reda A. A. Ammar, Abdulrahman Al-Warthan
(2012), Carbon nanotubes, science and technology part (I) structure, synthesis and characteri-
sation, 2004, Arabian Journal of Chemistry 5(1),1–23.
43. Anon, O. F., & Poole, C. P., Jr. (2003). Introduction to nanotechnology. Hoboken: Wiley.
44. Zahoor, M., Ahmad, M., Karim, S., Waheed, K., Ali, G., Hussain, S., Hussain, S., & Nisar,
A. (2018). Tungsten oxide multifunctional nanostructures: Enhanced environmental and
sensing applications. Materials Chemistry and Physics, 221. https://doi.org/10.1016/j.
matchemphys.2018.09.034.
22  Nanostructure Material-Based Sensors for Environmental Applications 587

45. Alexiadou, M., Kandyla, M., Mousdis, G., & Kompitsas, M. (2017). Pulsed laser deposition
of ZnO thin films decorated with Au and Pd nanoparticles with enhanced acetone sensing
performance. Applied Physics A, 123, 262.
46. Maduraiveeran, D.  G., & Jin, W. (2017). Nanomaterials based electrochemical sensor and
biosensor platforms for environmental applications. Trends in Environmental Analytical
Chemistry, 13, 10–23. https://doi.org/10.1016/j.teac.2017.02.001.
47. Pokropivny, V., Lõhmus, R., Nova, I., Pokropivny, A., & Vlassov, S. (2007). Introduction in
nanomaterials and nanotechnology.
48. Wei, Y., Yi, G., Xu, Y., Zhou, L., Wang, X., Cao, J., Sun, G., Chen, Z., Hari, B., & Zhang,
Z. (2017). Synthesis, characterization, and gas-sensing properties of Ag/SnO2/rGO compos-
ite by a hydrothermal method. Journal of Materials Science: Materials in Electronics, 28,
17049–17057. https://doi.org/10.1007/s10854-­017-­7630-­y.
49. Tharsika, T., Haseeb, A.  S. M.  A., Akbar, S.  A., Sabri, M.  F. M., & Hoong, W.  Y. (2014).
Enhanced ethanol gas sensing properties of SnO2-Core/ZnO-shell nanostructures. Sensors, 14,
14586–14600. https://doi.org/10.3390/s140814586.
50. Li, K., Diaz, D. C., He, Y., Campbell, J. C., Tsai, C., Li, K., Diaz, D. C., He, Y., & Campbell,
J. C. (1995). Electroluminescence from porous silicon with conducting polymer film contacts.
Electroluminescence film contacts from porous silicon with conducting polymer. Applied
Physics Letters, 2394, 1992–1995. https://doi.org/10.1063/1.111625.
51. Pramanik, S., Das, G., & Karak, N. (2013). Facile preparation of polyaniline nanofibers modi-
fied bentonite nanohybrid for gas sensor application. RSC Advances, 3, 4574–4581. https://doi.
org/10.1039/C3RA22557G.
52. Lee, J.  Y., Nguyen, T.  L., Park, J.  H., & Kim, B.-K. (2016). Electrochemical detection of
hydrazine using poly(dopamine)-modified electrodes. Sensors, 16. https://doi.org/10.3390/
s16050647.
53. Navale, S. T., Khuspe, G. D., Chougule, M. A., & Patil, V. B. (2014). Camphor sulfonic acid
doped PPy/α-Fe2O3 hybrid nanocomposites as NO2 sensors. RSC Advances, 4, 27998–28004.
https://doi.org/10.1039/C4RA02924K.
54. Wang, H., Zhang, Y., Ma, H., Ren, X., Wang, Y., Zhang, Y., & Wei, Q. (2016). Electrochemical
DNA probe for Hg(2+) detection based on a triple-helix DNA and Multistage Signal
Amplification Strategy. Biosensors and Bioelectronics, 86, 907–912. https://doi.org/10.1016/j.
bios.2016.07.098.
55. Abdul-wahab, S. A., Al-alawi, S. M., & El-zawahry, A. (2002). Patterns of SO 2 emissions: A
refinery case study. Environmental Modelling & Software, 17, 563–570.
56. Jiménez-Cadena, G., Riu, J., & Rius, F. X. (2007). Gas sensors based on nanostructured mate-
rials. The Analyst, 132, 1083–1099. https://doi.org/10.1039/B704562J.
57. Comini, E. (2006). Metal oxide nano-crystals for gas sensing. Analytica Chimica Acta, 568,
28–40. https://doi.org/10.1016/j.aca.2005.10.069.
58. Wang, Q., Liu, F., Lin, J., & Lu, G. (2016). Gas-sensing properties of In-Sn oxides compos-
ites synthesized by hydrothermal method. Sensors and Actuators B: Chemical, 234, 130–136.
https://doi.org/10.1016/j.snb.2016.04.042.
59. Verma, M. K., & Gupta, V. (2012). A highly sensitive SnO 2-CuO multilayered sensor struc-
ture for detection of H 2S gas. Sensors and Actuators B: Chemical, 166–167, 378–385. https://
doi.org/10.1016/j.snb.2012.02.076.
60. Basyooni, M. A., Shaban, M., & El Sayed, A. M. (2017). Enhanced gas sensing properties
of spin-coated Na-doped ZnO nanostructured films. Scientific Reports, 7, 41716. https://doi.
org/10.1038/srep41716.
61. Wu, J., Tao, K., Miao, J., & Norford, L. K. (2018). Three-dimensional hierarchical and super-
hydrophobic graphene gas sensor with good immunity to humidity. In 2018 IEEE Micro
Electro Mech. Syst. (pp. 901–904). https://doi.org/10.1109/MEMSYS.2018.8346702.
62. Keshtkar, S., Rashidi, A., Kooti, M., Askarieh, M., Pourhashem, S., Ghasemy, E., & Izadi,
N. (2018). A novel highly sensitive and selective H2S gas sensor at low temperatures based on
588 V. Srikanth et al.

SnO2 quantum dots-C60 nanohybrid: Experimental and theory study. Talanta, 188, 531–539.
https://doi.org/10.1016/j.talanta.2018.05.099.
63. Algadri, N. A., Hassan, Z., Ibrahim, K., & AL-Diabat, A. M. (2018). A high-sensitivity hydro-
gen gas sensor based on carbon nanotubes fabricated on glass substrate. Journal of Electronic
Materials, 47, 6671–6680. https://doi.org/10.1007/s11664-­018-­6537-­6.
64. Zhao, W., Yang, C., Zou, D., Sun, Z., & Ji, G. (2017). Possibility of gas sensor based on
C20 molecular devices. Physics Letters A, 381, 1825–1830. https://doi.org/10.1016/j.
physleta.2017.03.038.
65. Rahimi, R., Kamalinahad, S., & Solimannejad, M. (2018). Adsorption of rare gases on the
C20nanocage: A theoretical investigation. Materials Research Express, 5, 35006. https://doi.
org/10.1088/2053-­1591/aab0e3.
66. Hamouche, H., Makhlouf, S., Chaouchi, A., & Laghrouche, M. (2018). Humidity sensor based
on keratin bio polymer film. Sensors and Actuators A: Physical, 282, 132–141. https://doi.
org/10.1016/j.sna.2018.09.025.
67. Dubourg, G., Segkos, A., Katona, J., Radovic, M., Savic, S., Niarchos, G., Tsamis, C., &
Crnojević-Bengin, V. (2017). Fabrication and characterization of flexible and miniaturized
humidity sensors using screen-printed TiO2 nanoparticles as sensitive layer. Sensors, 17, 1854.
https://doi.org/10.3390/s17081854.
68. Leng, X., Luo, D., Xu, Z., & Wang, F. (2018). Modified graphene oxide/Nafion composite
humidity sensor and its linear response to the relative humidity. Sensors and Actuators B:
Chemical, 257, 372–381. https://doi.org/10.1016/j.snb.2017.10.174.
69. Ding, X., Chen, X., Chen, X., & Zhao, X. (2018). A QCM humidity sensor based on fullerene/
graphene oxide nanocomposites with high quality factor. Sensors and Actuators B: Chemical,
266. https://doi.org/10.1016/j.snb.2018.03.143.
70. Rodahl, M., Höök, F., & Kasemo, B. (1996). QCM operation in liquids: An explanation of
measured variations in frequency and Q factor with liquid conductivity. Analytical Chemistry,
68, 2219–2227. https://doi.org/10.1021/ac951203m.
71. Hens, S., Cunningham, G., McGuire, G., & Shenderova, O. (2011). Nanodiamond-assisted
dispersion of carbon nanotubes and hybrid nanocarbon-based composites. Nanoscience and
Nanotechnology Letters, 3, 75–82. https://doi.org/10.1166/nnl.2011.1123.
72. Li, X., Chen, X., Yao, Y., Li, N., Chen, X., & Bi, X. (2013). Multi-walled carbon nanotubes/
graphene oxide composites for humidity sensing. IEEE Sensors Journal, 13, 4749–4756.
https://doi.org/10.1109/JSEN.2013.2273615.
73. Saha, D., & Das, S. (2018). Development of fullerene modified metal oxide thick films for
moisture sensing application. Materials Today: Proceedings, 5, 9817–9825. https://doi.
org/10.1016/j.matpr.2017.10.172.
74. Drummond, T. G., Hill, M. G., & Barton, J. K. (2003). Electrochemical DNA sensors. Nature
Biotechnology, 21, 1192–1199. https://doi.org/10.1038/nbt873.
75. Lei, W., Si, W., Xu, Y., Gu, Z., & Hao, Q. (2014). Conducting polymer composites with gra-
phene for use in chemical sensors and biosensors. Microchimica Acta, 181, 707–722. https://
doi.org/10.1007/s00604-­014-­1160-­6.
76. Ahmad, Z., Shah, A., Siddiq, M., & Kraatz, H.-B. (2014). Polymeric micelles as drug delivery
vehicles. RSC Advances, 4, 17028–17038. https://doi.org/10.1039/C3RA47370H.
77. Dai, H., Wang, N., Wang, D., Ma, H., & Lin, M. (2016). An electrochemical sensor based
on phytic acid functionalized polypyrrole/graphene oxide nanocomposites for simultaneous
determination of Cd(II) and Pb(II). Chemical Engineering Journal, 299, 150–155. https://doi.
org/10.1016/j.cej.2016.04.083.
78. Deng, L., Zhou, Z., Li, J., Li, T., & Dong, S. (2011). Fluorescent silver nanoclusters in hybrid-
ized DNA duplexes for the turn-on detection of Hg2+ ions. Chemical Communications, 47,
11065–11067. https://doi.org/10.1039/C1CC14012D.
79. Farias, J. S., Zanin, H., Caldas, A. S., dos Santos, C. C., Damos, F. S., & de Cássia Silva Luz,
R. (2017). Functionalized multiwalled carbon nanotube electrochemical sensor for determina-
tion of anticancer drug flutamide. Journal of Electronic Materials, 46, 5619–5628. https://doi.
org/10.1007/s11664-­017-­5630-­6.
22  Nanostructure Material-Based Sensors for Environmental Applications 589

80. Jeong, H., Nguyen, D. M., Lee, M. S., Kim, H. G., Ko, S. C., & Kwac, L. K. (2018). N-doped
graphene-carbon nanotube hybrid networks attaching with gold nanoparticles for glucose non-­
enzymatic sensor. Materials Science and Engineering: C, 90, 38–45. https://doi.org/10.1016/j.
msec.2018.04.039.
81. Liu, Y., Kannegulla, A., Wu, B., & Cheng, L.-J. (2018). Quantum dot fullerene-based

molecular beacon nanosensors for rapid, highly sensitive nucleic acid detection. ACS Applied
Materials & Interfaces, 10, 18524–18531. https://doi.org/10.1021/acsami.8b03552.
Chapter 23
Nanostructured MoS2 as Non-noble
Metal-­Based Cocatalyst for Photocatalytic
Applications

Murthy Muniyappa, Manjunath Shetty, Mahesh Shastri,


S. Jagadeesh Babu, M. Navya Rani, Prasanna D. Shivaramu,
and Dinesh Rangappa

23.1  Introduction

A chemical reaction takes place by the influence of light or energy packets (pho-
tons) known as photoreaction. Photoreaction can be accelerated by a catalyst; hence
the name photocatalysis came in to exist. In natural photosynthesis plants prepare
their own food by consuming light and CO2; likewise artificial photocatalysis has
developed for various applications. In 1972, Fujishima et al. first time reported the
photoelectrochemical splitting of water by semiconducting TiO2 [1]. Thereafter,
TiO2 has been studied and used as a prominent catalyst for environmental remediation
applications such as degradation of organic dyes, H2 generation by water splitting,
heavy metal reduction, CO2 reduction, etc. In the photocatalysis process, light with
appropriate energy equivalent to the bandgap of a semiconductor excites free
electrons from valence band to conduction band and creates electron and hole pairs.
The generated electrons and holes involved in a redox reaction to form reactive
oxygen species (ROS); thus formed ROS will degrade the pollutants. TiO2 is the
most studied semiconducting material with a wide bandgap which requires UV light
for an electron to excite from valence band to conduction band. In solar spectrum,
only 5% of spectrum consists of UV light, and more than 40% includes visible light;
therefore visible light-driven photocatalyst (less bandgap) has been explored for
photocatalytic applications. For shifting bandgap from the UV region to the visible
region, many efforts have been made such as doping, surface modification, and
synthesis of binary and ternary composites. Recombination and conductivity of

M. Muniyappa · M. Shetty · M. Shastri · S. Jagadeesh Babu · P. D. Shivaramu · D. Rangappa (*)


Department of Applied Sciences, Visvesvaraya Technological University, Center for
Postgraduate Studies, Muddenahalli, Chikkaballapur, Karnataka, India
M. Navya Rani
School of Basic and Applied Sciences, Dayanand Sagar University,
Bengaluru, Karnataka, India

© Springer Nature Switzerland AG 2021 591


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6_23
592 M. Muniyappa et al.

charge carriers are the major drawbacks in wide bandgap materials. To solve these
problems, noble metals are applied as cocatalysts.
Noble metals like Pt, Au, and Pd [2] act as cocatalysts for various semiconduct-
ing metal oxides, sulfides, and carbonitrides for photocatalytic applications.
Cocatalysts act as an electron sink that significantly enhances the photocatalytic
efficiency by suppressing recombination of charge carriers. Cocatalysts also
increase proton reduction by providing large active sites with increased conductiv-
ity of charge carriers as shown in Fig. 23.1. By loading the noble metals as cocata-
lysts, photocatalytic efficiency can be improved. However, it is not economically
feasible because of the high cost. To avoid these noble metals, transition metal sul-
fides like MoS2, NiS, ZnS, etc. have been loaded as cocatalysts because of high
abundance, low cost, and excellent optical and electrical properties. Among the
various metal dichalcogenides, MoS2 has attracted considerable interest owing to its
significant advantages over noble metals because of its low cost, nontoxicity, and
highly active and excellent chemical stability.
Molybdenum disulfide is made up of two-dimensional sheets binding together
by van der Waals force as in the case of graphene. In almost all the fields of science
and technology, MoS2 has its applications. However, MoS2 has an advantage over
other materials because of its low cost, abundance, non-standard shape, and
adjustable bandgap with excellent optical properties. Therefore, it attracts a lot of
attention worldwide.
This chapter mainly focuses on the study of crystal structures, basic properties of
MoS2, and synthesis strategies like solvo-/hydrothermal and chemical exfoliation
methods. More importantly, recent studies on non-noble metal-based MoS2-loaded
semiconductors for hydrogen generation and strategies for enhanced H2 generation
are explored. Other applications of MoS2 as non-noble metal-based cocatalysts such
as photocatalytic dye degradation, toxic metal ion reduction, bacterial disinfection,
and mechanisms are discussed.

Fig. 23.1  Schematic of


cocatalyst-loaded
semiconductor for H2
generation
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 593

23.2  Crystal Structure and Basic Properties of MoS2

MOS2 is one of the transition metal dichalcogenide generally denoted as MX2 where
M represents the transition Mo metal and X is the S chalcogen. MoS2 has a layered
structure as shown in Fig. 23.2a. In pristine form, MoS2 layers bind with weak van
der Waals forces, analogous to graphite structure. One sheet of MoS2 consists of
three atomic layers of Mo and S in the form of S-Mo-S. Each Mo atomic layer is
sandwiched between two S layers where Mo acts as the backbone. The distance
between the consecutive Mo atomic layer is found to be 6.5 Å. Depending on crystal
structure MoS2 can be classified into three types, namely, 2H, 3R, and 1T. As shown
in Fig. 23.2b, in 2H phase two trigonal prismatic crystal structures form a hexagonal
symmetry, but in 3R-type crystal structure, three layers of trigonal prismatic crystal
structure form a rhombohedral symmetry, whereas in 1T-type tetragonal crystal
structure consists of Mo atoms at every alternative site without any symmetry that
can be seen in Fig. 23.2b. Among all three crystal structures, 2H type is the most
common and stable 2D structure, and it can be easily obtained by liquid exfoliation,
solvothermal, and scotch tape method.
The electrical properties of the MoS2 by standard DFT calculations is shown in
Fig. 23.3a. Monolayer MoS2 shows a direct bandgap of 1.8 eV, whereas bulk MoS2
shows 1.21 eV. The shift in the bandgap is due to the quantum confinement effect
and hybridization of Mo 4dz and S 3pz orbitals. When layers of MoS2 decreases,
strong interlayer coupling effect comes into picture. Due to out-of-plane orbitals,
transition states occur at Ѓ point which results in a shift in the bandgap as shown in
Fig. 23.3a. The conduction band edge potential is depending upon the number of
layers as shown in Fig. 23.3b. In bulk MoS2 the CB edge position is at −0.16 eV
which is higher than the H+ ion reduction. In case of monolayer MoS2, the conduction
band edge shifts to −0.53  eV, which is a more negative reduction potential.
Therefore, bulk MoS2 is not suitable for catalytic applications. For example, in TiO2
conduction band, the edge is at −0.50  eV, which is within the water splitting
potential, but for monolayer MoS2, the conduction band potential is −0.53 eV that

Fig. 23.2 (a) Three-dimensional structure of single-layer MoS2 and (b) different structures of
MoS2. Reprinted with permission from [3, 4], nature publishing group
594 M. Muniyappa et al.

Fig. 23.3 (a) Monolayer band structure of MoS2. (b) Size-dependent variation in bandgap of
MoS2. Reprinted with permission from [4]. Copyright 2012, nature publishing group

is more negative than the TiO2, hence impossible to load single-layer MoS2 because
of more negative (−0.53 eV) conduction edge potential than TiO2 (0.50 eV). A few
layer MoS2 can have less than −0.50 eV which can be loaded as a cocatalyst for
TiO2. Always the conduction band edge potential should be more negative than the
H+ ion reduction potential to produce H2 gas.
Semiconductor MoS2 (2H) do not exhibit high performance due to the slow rate
of charge transfer and lack of active centers in electrocatalyst and photocatalysts
[5]. The electrical properties of MoS2 can be tuned by changing the crystal structure/
phase from 2H to 1T phase [6]. 1T phase provides less charge transport resistance
as well as a high charge transfer rate during the photocatalysis process. In addition,
metallic-phase MoS2 provides a large number of active sites than the semiconducting
phase. Therefore, conversion of semiconducting to metallic MoS2 is a great success,
and an increase in catalytic activity can be easily achieved. The metallic-phase
MoS2 in catalysis, related to energy conversion, is of great importance for use in
practical environmental remediation applications.

23.3  S
 ynthesis of Noble Metal-Free
MoS2-Based Nanocomposites

23.3.1  Hydrothermal/Solvothermal Method

Hydrothermal or solvothermal synthesis methods are the most studied and simple
techniques to synthesize various oxides as well as sulfides. In these methods, the
precursors are dissolved in water or organic solvent and kept in a sealed vessel or
autoclave at high pressure and temperature usually above the boiling point of the
solvents, ranging from 60 to 250  °C.  The hydrothermal/solvothermal method
facilitates the precise control over the particle size, and morphology can be
controlled by varying the reaction parameters such as reaction time, temperature,
pressure, pH, surfactants, or capping agents. There have been many metal o­ xides/
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 595

MoS2, metal sulfides/MoS2, and carbonitrides/MoS2 composites synthesized


through the solvothermal/hydrothermal method as shown in Table 23.1.
For example, MoS2 nanosheet-coated TiO2 nanobelts can be synthesized by
hydrothermal method [7]. In a synthesis process, sodium molybdate (30 mg) and
thioacetamide (60  mg) act as precursors of Mo and S, and presynthesized TiO2
nanobelts are dispersed with the precursors, then dissolved in water, and placed in
hydrothermal autoclave at 200 °C for 24 h. At subcritical temperature, the positive
and negative ions will get saturated and started to nucleate, growth in particular two
directional plane  to form sheet like morphology, At the  surface or walls of the
autoclave. After completing the reaction autoclave naturally cooled down to room
temperature to obtain a black product and it is washed to remove dissolved
contaminants and other dissolved impurities. Further, dried overnight at 50 °C. The
synthesized black powder is characterized to confirm MOS2 nanosheets coated on
TiO2 nanobelts as shown in Fig. 23.4.

23.3.2  Chemical Exfoliation

Layered bulk materials can be exfoliated into single- or few-layered materials by


organic solvents which is known as chemical exfoliation. In 1986, for the first time,
Per Joensen et al. reported chemical exfoliation of single-layer MoS2 by lithium and
water intercalation [16]. In the synthesis process, bulk MoS2 (325 mesh) is soaked

Table 23.1 Noble metal-free MoS2-based composites synthesized through solvothermal/


hydrothermal method
Morphology of MoS2
Sample name Solvent Parameters (cocatalyst) Refs.
TiO2 @ MoS2 Water 160 °C, Nanosheet [7]
24 h
g-C3N4-MoS2 Water 160 °C, Wrinkles [8]
24 h
CdS/MoS2 Water + 180 °C, Nanosheets [9]
l-cysteine 12 h
MoS2-ZnO Water 210 °C, Nanosheet [10]
24 h
ZnTCPP-MoS2/ZnO Water 200 °C, Nanosheet [11]
24 h
TiO2/MoS2/graphene Water/ethanol 210 °C, Nanosheet [12]
composite 24 h
MoS2-g-C3N4 Water 220 °C, Nanosheet [13]
24 h
MoS2/graphene/CdS Water + 120 °C, Nanosheet [14]
l-cysteine 24 h
MoS2/SrZrO3 Water 200 °C, Nanoflower [15]
24 h
596 M. Muniyappa et al.

Fig. 23.4 (a) FE-SEM images of TiO2@ MoS2 and (b) HR-TEM images of TiO2@ MoS2 synthe-
sized by hydrothermal method. Reprinted with permission from [7]. Copyright © 2013 Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim

in 1.6 M n-butyl lithium in hexane solvent for 48 h with Ar atmosphere in a dry box,
then washed with hexane, and immersed in water followed by sonication. When
intercalated Li ions come into contact with water, H2 gas evolution takes place.
Expansion of H2 gas leads to MoS2 layer exfoliation by breaking weak van der
Waals force between the layers and separating into few layers of MoS2. The extent
of Li ion intercalation deep into the crystalline structure is responsible for the
number of layers to be separated. Therefore, Li ion intercalation is very important
in separating the layers and also the thickness of the exfoliated layers. Various
solvents such as methanol, ethanol, and isopropanol have been used to exfoliate the
layered metal sulfides.
CdSe/MoS2 nanocomposite can be synthesized by wet chemical process-assisted
liquid exfoliation method. In this process CdSe nanobelts synthesized by a simple
wet chemical technique where cadmium cation and selenium anion reaction takes
place (Lewis acid-base reaction) under suitable surfactant such as octylamine [17].
Then, Cd and Se atoms start to nucleate and started to grow into CdSe nanostructures.
MoS2 is synthesized by the liquid exfoliation method by intercalating 1.6 M n-butyl
lithium in hexane followed by drying and sonication in water. The intercalated Li
ions reacts with water and liberate H2 gas. As liberated H2 gas while expansion tends
to exfoliate the layers of MoS2. The obtained MoS2 paste is sonicated in water for a
while until full dispersion. The dispersed MoS2 solution was added into the CdSe
solution and stirred for half an hour to complete the coating of MoS2.
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 597

23.3.3  Ball Milling

Ball milling is one type of top-down approach, where bulk materials are chopped
into nanosized materials by grinding in a jar by high-speed rotating balls. It has been
considered as an eco-friendly, high-yield, simple technique for synthesizing solid-­
state composite materials with pure form.
The MoS2-fullerene nanocomposite is synthesized by ball milling method. The
MoS2 bulk powder and fullerene powder are placed in a planetary high-speed ball
milling jar containing 3-mm-diameter stainless steel balls. Then, the planetary ball
milling is set to rotate at a speed of 500 rpm over the time period of 48 h [18]. When
rotating at a high speed, the force exerted by balls breaks down the interlayer van
der Waals force between the layers of bulk MoS2 and forms thin nanosheets. The π
bonding between carbon atoms in the fullerene will break down and bond with
MoS2 nanosheets. The size of the particles depends upon the rotation speed, time,
and size of the balls. The ball milling process consumes more time to form nano
particles.

23.4  Applications of Nanostructured MoS2

23.4.1  M
 oS2 as Non-noble Metal-Based Cocatalyst
for Hydrogen Generation

Photocatalytic hydrogen generation by water splitting is one of the best techniques


for mitigating energy problems, across the world. Many semiconducting materials
have been employed as photocatalysts for H2 generation. For better performance,
photocatalyst should have long-range absorption capacity, suitable band position,
and high stability, and it should be low cost. Various materials such as metal
phosphides, sulfides, and carbonitrides have been used in H2 generation under
natural solar light as well as artificial light. Among these materials, MoS2 shows
significant catalytic activity because of its unique 2D structure and optical and
electrical properties.
In the photocatalytic H2 generation mechanism, the electrons at the conduction
band absorbs the appropriate amount of light energy and get excited to valance
band. The excited electrons from valance band to the conduction band reduce H+
ions into H2 at the conduction band. The overall water splitting reaction is as follows:

H 2 O → H 2 + 1 / 2O2

Once the generated electrons recombine with holes due to recombination, it is


wasted because electron will not be able to participate in reducing H+ ion. To avoid
electron-hole recombination, cocatalyst has been loaded to the semiconductor. A
cocatalyst not only reduces the recombination of charge carriers, but it also helps
598 M. Muniyappa et al.

to  generate more reactive sites in semiconductor; thus enhanced photocatalytic


activity can be achieved. For example, pristine TiO2 shows less photocatalytic
activity due to the inefficient use of electrons for reducing H+ ions into H2. For
efficient use of generated electrons to participate in the water splitting reaction,
MoS2 has loaded into TiO2 as cocatalyst. The photocatalytic H2 generation was
found to be 150.7μmol h−1 g−1, which is 48.6 times higher than the activity of TiO2
alone. The schematic diagram of TiO2/MoS2 nanocomposite for H2 generation is
shown in Fig. 23.5.
MoS2 has been loaded as cocatalyst for various metal oxides, metal sulfides,
carbonitrides, etc. as shown in Table 23.2.

Strategies to Improve Hydrogen Generation

MoS2 as non-noble metal-based cocatalyst being loaded for oxides, sulfides, and
carbonitrides for improving photocatalytic H2 evolution. To overcome problems
such as low conductivity, less absorption capacity of solar light, electron-hole
recombination, and a smaller number of active sites, numerous strategies have been
applied for enhanced H2 generation. Some of the strategies to improve photocatalytic
non-noble metal-based H2 generation are by usage of heterojunction composite
materials, converting semiconducting phase to metallic phase, controlling
morphology, Z-scheme-based photocatalysis, and synthesizing composites with
carbon-based materials that are discussed in the next section.

Fig. 23.5 Schematic
diagram of TiO2/MoS2
non-noble metal-based
nanocomposite for
photocatalytic water
splitting. Adopted from
and reproduced from [19]
with permission from the
Royal Society of
Chemistry
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 599

Table 23.2  List of noble metal-free MoS2 as cocatalyst for photocatalytic H2 generation
Light source and H2 yield
Catalyst Sacrificial agent duration (mmol h−1 g−1) Refs.
4% MoS2-TiO2 Methanol 300 W xenon 150.7μmol h−1 [12]
lamp,6 h
50% MoS2-TiO2 0.35 M Na2S and 0.25 M Na2 300 W xenon 1.6 mmol h−1 g−1 [7]
SO3 lamp
MoS2@Cu2O 25%(v/v) methanol + Na2SO4 350 W xenon 230μmol h−1 [20]
(0.1 mol L−1, pH = 7) lamp,6 h
1.00 wt% 1 M Na2S and 1 M Na2SO3 300 W xenon 768μmol h−1 g−1 [10]
MoS2-ZnO lamp
3 wt% MoS2@ 0.35 M Na2S/0.25 M Na2SO3 300 W xenon 0.42 mmol h−1 g–1 [21]
ZnxCd1−xS lamp
0.6 wt% MoS2/ 0.5 M Na2SO3/0.43 M Na2S 300 W xenon 153μmol h−1 [22]
ZnIn2S4 lamp
0.05 wt% 0.35 M Na2S/0.25 M Na2SO3 300 W mercury 5.31 mmol h−1 [15]
MoS2/SrZrO3 vapor lamp
0.5 wt% 0.1 M Na2SO3 Xe arc lamp 23.1μmol h−1 [13]
MoS2/g-C3N4
0.5 wt% 0.1 M Na2S/0.1 M Na2SO3 300 W xenon 23.1μmol h−1 [17]
CdSe-MoS2 lamp

Synthesizing Heterojunction Nanocomposites

The heterojunction is a transition of a band positions of a composite formed by a


combination of two different semiconductors. These heterogeneous structures are
formed based on the bandgap, electron affinity, and band position to form new
energy levels. There are three types of heterojunction semiconductors, namely, type
I, type II, and type III. In type I the band positions of one material is overlapped with
another semiconductor. The band position of one semiconductor is higher than
another called as type II semiconductors. In type III, the bandgaps are not overlapped
or broken bandgap, attained by the combination of semimetal and semiconductor.
In 2020, Wei Zhao synthesized CdS/MoS2 binary composites for photocatalytic
hydrogen generation without noble metals as cocatalyst [23]. The CdS nanowire-­
coated MoS2 nanosheets shows 1.79 mmol g−1 h−1 hydrogen generation, which is
~36 times higher catalytic activity than bare CdS.
The CdS/MoS2 nanocomposite forms a heterojunction with different conduction
band edge potentials. The conduction band of CdS is lesser than MoS2, which helps
trap the excited electrons and utilize for the reduction of H+ ion to hydrogen by
suppressing electron-hole recombination as shown in Fig.  23.6. Formation of a
heterojunction is the key to enhance photocatalytic activity.
The MoS2-CdS-TaON nanocomposite shows enhanced hydrogen generation
[24], where MoS2 and CdS act as a cocatalyst for photocatalytic hydrogen generation
under visible light irradiation. The composite shows 628.5μmol  h−1 hydrogen
generation which is ~70 times higher activity than bare TaON with Pt as cocatalyst.
This confirmed the importance of heterojunction formation by CdS nanocrystals
600 M. Muniyappa et al.

H2
-1.0 2H+
Conduction band
2.4eV
0 E(H+/H2)
hυ 1.7eV
Valence band
1.0
TEOA

CdS MoS2
Oxidation
product

Fig. 23.6  Schematic diagram of non-noble metal-based CdS/MoS2 nanocomposite for photocata-
lytic water splitting. Reprinted from [23]. Copyright 2020, with permission from Elsevier

and TaON nanospheres. The excited electrons from CdS easily fall into the
conduction band of TaON because of less negative conduction band position. Then,
the electrons are easily trapped by MoS2 nanocrystals before recombining with
holes, which suppress the electron-hole recombination efficiently. Loading of MoS2
into CdS/TaON nanocomposite not only reduces charge recombination but also
reduces the photo corrosion of CdS nanocrystal. Therefore, the photocatalytic
hydrogen generation of nanocomposite MoS2-CdS-TaON is increased. The
formation of heterojunctions by binary and ternary compounds shows substantially
very high photocatalytic hydrogen generation.

Phase Engineering in MoS2

In general, MoS2 exists in three polymorphs, namely, 2H, 3R, and 1T phases, as
discussed in the introduction part. The electronic properties of the materials depend
upon the phase of the material. In 2H phase, d orbitals are completely filled, whereas
in 1T phase, d orbitals are filled with unpaired electrons, and the presence of empty
doubly degenerate orbitals makes it metallic [25].
A few-layered MoS2 (2H phase) shows a very negligible amount of hydrogen
generation (0.05 mmol h−1 g−1) because of its inability to transfer free electrons for
reducing H+ ion to H2. Normally, MoS2 is loaded into conductive carbonaceous
materials like RGO, nitrogen-doped RGO, CNT, etc., to enhance electron transfer
rate and suppression of charge recombination. Still, the conductivity of the MoS2 is
the key issue for enhanced water splitting process. Therefore, U.  Mitra et  al.
synthesized 1T-phase metallic MoS2 by exfoliation method through the Li ion
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 601

intercalation technique and tested the photocatalytic activity of prepared 1T-phase


MoS2 under 100 W halogen lamp [25]. The photocatalytic activity was found to be
30 mmol h−1 g−1 which is 600-fold higher than the 2H-phase MoS2. In the 1T phase,
the partially filled d orbital exhibits metallic nature which influences the conductivity
of electrons with enhanced charge transfer rate. In MoS2, sulfur edges are more
reactive compared to the basal plane of Mo atoms. The S edges in 1T phase are more
exposed for catalytic reaction compared to 2H phase because of its octahedral
coordination bonding with Mo atoms. Therefore, the metallic-phase MoS2 provides
the large number of reactive sites for water splitting.
Metallic MoS2 is loaded as cocatalyst with other semiconducting materials like
g-C3N4, CdS, TiO2, and other organic dye molecules for increased light absorption
range as well as the absorption capacity of the photocatalyst. Therefore, visible
light-driven photocatalytic water splitting can be achieved with enhanced water
splitting efficiency. Hence, phase-engineered MoS2 loaded as cocatalyst with other
semiconducting materials increases water splitting efficiency.

Z-Scheme-Based MoS2 Nanocomposites

Nowadays, solar light-driven photocatalytic water splitting is the most common and
viable technique for H2 generation. For water splitting the band position of both
valence band and conduction band is very important. The conduction band should
be more negative than H+ reduction potential (0 V vs NHE) and more positive than
the oxidation potential of water (1.23 vs NHE). For achieving a suitable band edge
position and charge recombination suppression, single-material semiconductors are
not suitable. However, in most of the water splitting process, sacrificial agents are
used to trap generated holes. To generate H2 in pure water, without sacrificial agents,
is a challenging task. Solar light-driven photocatalytic water splitting in pure water
can be achieved by Z-scheme charge transmission process.
In Z-scheme system, there will be two catalysts combined to form one oxidation
photocatalyst and another reduction photocatalyst. In this kind of mechanism, there
will be a greater number of electrons in the reduction catalyst, which helps reduce
more H+ ions, whereas holes in the oxidation photocatalyst will help oxidize surface-­
adsorbed water molecules.
Recently, D. Wei et al. studied Z-scheme non-noble metal-based MoS2/CdS-WO3/
MnO2 nanocomposite for photocatalytic H2 generation under solar light [26]. Here,
CdS nanoparticles are grown on WO3 nanorods by precipitation method followed
by MoS2 and MnO2 coating by photoreduction and photooxidation process,
respectively. The photocatalytic activity is carried out under 300  W xenon lamp
with cutoff filters in visible range without any cocatalyst. Various wt% loading
combinations of MoS2 and MnO2 into CdS-WO3 composite are tested for
photocatalytic H2 generation. Only 2 wt%-loaded MoS2 and 1 wt%-loaded MnO2
show the highest catalytic activity of 0.76μ moles of hydrogen and 0.24μ moles of
oxygen production with 12 h of light irradiation without any sacrificial agents.
602 M. Muniyappa et al.

The mechanism of ZCS-charge transfer process in the MoS2/CdS-WO3/MnO2


nanocomposite is shown in Fig. 23.7. The valance band position of WO3 is at 2.8 eV
vs NHE; therefore WO3 is acting as an oxidation photocatalyst. Further, the
conduction band position of CdS is at −0.51 eV vs NHE, which is a more negative
potential than WO3 and H+ reduction potential; therefore it acts as a reduction
catalyst. When the light is irradiated to the synthesized MoS2/CdS-WO3/MnO2
photocatalyst, the electrons are excited into the conduction band leaving holes at the
valance band of both WO3 and CdS semiconductors. The holes generated at valance
band of WO3 (more positive than water oxidation potential of 1.23 eV vs NHE) will
be captured by the MnO2 cocatalyst without allowing holes to recombine with
electrons. The captured holes at MnO2 will help oxidize surface-adsorbed water for
O2 evolution. The excited electrons from the valance band of CdS will be captured
by the loaded MoS2 cocatalyst, before recombination with holes. The electrons will
actively participate in the reduction of H+ ions more efficiently. Hence, the
photocatalytic activity of the prepared MoS2/CdS-WO3/MnO2 nanocomposite
shows the highest photocatalytic activity without any sacrificial agent under
visible light.
Many semiconducting nanocomposites like zinc stannate-based materials with
MoS2 cocatalyst through ZCS-charge transfer process show enhanced photocatalytic
activity because of its excellent charge separation and visible light absorption
capacity. Under solar light irradiation, both O2 and H2 generations can be achieved
by combining two semiconductors with MoS2 as cocatalyst without any sacrificial
agents through the Z-scheme charge transfer process.

Fig. 23.7  The mechanism


of ZCS-charge transfer
process in the MoS2/
CdS-WO3/MnO2
nanocomposite. Reprinted
from [26]. Copyright 2020,
with permission from
Elsevier
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 603

Morphology Control

Certain physical properties of the cocatalyst, such as shape and structure, also mat-
ter for enhanced photocatalytic hydrogen generation. In most of the cases, MoS2
nanosheets are loaded as cocatalyst with metal oxides, sulfides, and carbonitrides
for water splitting reaction without any noble metals. Basically, MoS2 is a layered
structure with X-M-X structure, where X represents the sulfur atoms and M
represents the Mo atoms. One layer of Mo atoms is sandwiched between two layers
of sulfur atoms. For enhanced photocatalytic activity, there should be more reactive
sites for hydrogen reduction. Therefore, edge-rich MoS2 contains a greater number
of sulfur atoms, which act as reactive sites for hydrogen ion reduction. Monolayers
of MoS2 contain more sulfur edges compared to few-layered MoS2 sheets. But still
the efficiency of the photocatalytic activity of monolayered MoS2 shows less
compared with nanodots of MoS2. Therefore, morphology is also playing an
important role here.
In 2018, X. Shi et al. reported that MoS2 nanodots are loaded as cocatalysts with
g-C3N4 for photocatalytic water splitting process by photodeposition method [27].
The prepared g-C3N4/MoS2 nanodots exhibit an enhanced H2 generation of
600μmol  h−1  g−1, which is almost eight times higher than the monolayer MoS2-­
loaded g-C3N4. In MoS2 nanodots, sulfur atoms comprise of rich edges in comparison
to a monolayer of MoS2. Spherical morphology of the loaded MoS2 induces a
greater number of unsaturated sulfur atoms at the surface of the MoS2, thereby
increasing the number of reactive sites for H+ ion reduction. Due to the nanosized
effect of MoS2, the quantum confinement effect results in a wider bandgap. As a
result, the photo-absorption range of the C3N4/MoS2 nanodots is enhanced.
Morphology is also an important parameter in controlling the catalytic activity of
the photocatalysts. Compared to sheet or 2D morphology, spherical- or nanodot-­
loaded semiconductors are more beneficial for enhanced H2 generation without
expensive noble metals.

23.4.2  M
 oS2 as Non-noble Metal-Based Cocatalyst
for Dye Degradation

Photocatalytic dye degradation is an effective technique for environmental remedia-


tion applications. In an effort to develop low-cost, efficient, and eco-friendly photo-
catalysts, natural sun light-driven photocatalysis is an attractive technique. There
have been reports on various metal oxides and sulfides that are used as photocata-
lyst. Due to the wide bandgap, high charge recombination and lack of conductivity
cocatalyst are loaded to overcome these limitations. Many noble metals are used as
cocatalyst, due to the scarcity of high cost, and researchers are looking for cheaper
and efficient cocatalysts. MoS2 is one of the highly active and abundant materials.
Therefore, it is used as an alternative cocatalyst for dye degradation.
604 M. Muniyappa et al.

In photocatalytic dye degradation, upon irradiation of light, hydroxyl and super


oxide radicals are formed by redox reaction of electrons and holes. These generated
highly reactive radicals degrade the complex dye molecules into smaller molecules
of CO2 and water. The loaded MoS2 traps the excited electrons and transfers quickly
to participate in redox reaction, thereby suppressing the charge recombination as
shown in Fig.  23.9. Many semiconducting materials are loaded with MoS2 for
photocatalytic dye degradation under solar light.
For improved photocatalytic degradation, S.  Kumar et  al. synthesized ZnO-­
MoS2-­RGO nanocomposite for photocatalytic degradation of methylene blue dye
under irradiation of sunlight [28]. ZnO is the wide bandgap material, which is a UV
light-active material, for increasing the light absorption range in visible light. MoS2
was loaded for shifting the bandgap from UV region to visible region. By absorbing
light of energy equivalent to the bandgap of ZnO and MoS2, electrons are excited to
conduction band from the valance band. Then, the excited electrons were captured
by reduced graphene oxide (RGO) because the conduction band edge (−0.08 eV) is
less negative than MoS2 conduction band (−0.13 eV) edge potential. Electrons from
RGO reduce the H+ ions into H2O2, which is further oxidized into reactive hydroxyl
radicals. The generated hydroxyl radicals degrade the organic pollutants into smaller
byproducts as shown below. The schematic of the photocatalytic degradation
mechanism is shown in Fig. 23.8.

ZnO − MoS2 − RGO( catalyst ) + sunlight → e − ( CB) + h + ( VB)

e − ( ZnOCB) → e − ( MoS2 CB) → e − RGO

e − RGO + 2H + + O2 → H 2 O2

Fig. 23.8  The mechanism of ZnO-MoS2-RGO nanocomposite for the degradation of organic
dyes. Reprinted from [27], with permission from the Royal Society of Chemistry
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 605

H 2 O2 + e − + light → ∗OH + OH −

h + ( VB) + H 2 O → H + + ∗OH

∗OH + Organicdye → byproducts

Overall, there are three main factors that enhance the photocatalytic dye degrada-
tion. The factors are, improved light harvesting capacity both in UV and visible
region, suppression of charge carrier’s recombination by quickly transferring
excited electrons to RGO followed by generation of hydroxyl and super oxide radi-
cles and increased surface area, which adsorb large quantity of dye in turn increas-
ing the dye degradation.

23.4.3  M
 oS2 as Non-noble Metal-Based Cocatalyst
for Reduction of Toxic Metals

Water pollution is one of the major concerns all over the world. Industrial wastes
that contain organic dyes, chemicals, pharmaceutical chemicals, and heavy metals
are the main sources for water pollution. Heavy metals such as Cr(VI), Hg(II),
Fe(III), and Ni(II) uptake of these heavy metals result in severe health issues.
Compared to all other heavy metals, Cr(VI) shows heavy toxicity and is very
difficult to treat because it easily dissolves in water; therefore reduction of Cr(VI) to
less toxic Cr(III) is essential.
Heavy metal reduction by photocatalysis is a simple, easy, and more explored
technique. It involves excitation of e− from the VB to the CB by irradiating light of
wavelength equivalent to the bandgap of semiconductor. The excited electrons
reduce chromate ions into less toxic Cr3+ ions. The heavy metal reduction by
photocatalysis has some limitations such as electron-hole recombination, less
conductivity, and limited light absorption capacity. To overcome these limitations,
MoS2 has loaded as catalyst/cocatalyst for enhanced reduction of heavy metals by
photocatalysis process.
Recently in 2020, W. Pudkon et al. studied the ZnIn2S4/MoS2 synthesized by
biomolecule-assisted microwave method for the Cr(VI) reduction [29]. The reac-
tion mechanism of heavy metal reduction is shown in Fig. 23.9. When visible light
irradiated on ZnIn2S4/MoS2 composite, electrons from the valence band jump to
the conduction band leaving holes at the valance band. Depending on the valance
band position and potential difference, holes start to migrate from the more posi-
tive band position of MoS2 (1.89  eV) to a lower positive band position ZnIn2S4
(1.81 eV) and oxidize the water (1.23 eV vs NHE). However, the electrons from
the conduction band of the ZnIn2S4 (−1 eV) transfer to the less negative potential
edge of MoS2 (−0.07 eV) and reduce the chromate ions. The reactions involved are
as follows:
606 M. Muniyappa et al.

Fig. 23.9  The mechanism of ZnIn2S4/MoS2 for the reduction of Cr(VI) reduction. Reprinted from
[29], with permission from the Royal Society of Chemistry

ZnIn 2 S4 / MoS2 ( catalyst ) + sunlight → ZnIn 2 S4 / MoS2 e − ( CB) + h + ( VB) 


H 2 O + h + ( VB) ZnIn 2 S4 → O2 + 4H +

In acidic medium,

Cr2 O7 2 − + 14 H + + MoS2 6 e − ( CB) → 2 Cr 3+ + 7H 2 O

HCrO 4 − + 7H + + MoS2 3e − ( CB) → Cr 3+ + 4H 2 O

In basic medium,

CrO 4 2 − + 4H 2 O + MoS2 3e − ( CB) → Cr ( OH )3 + 5OH −



Further, to increase photocatalytic heavy metal reduction, Z-scheme-based mate-
rials have attracted researchers. In 2020 F. Mu et al. synthesized that polycarboxylic
group metal oxide framework incorporated (UiO-66-(COOH)2)ZnIn2S4/MoS2 com-
posite synthesized by hydrothermal method [30], where MoS2 acts as a cocatalyst
which enhances the charge separation as well as charge transfer process. Under
visible light irradiation, Cr(VI) reduced with a rate constant of 0.0685  min−1.
Therefore, polycarboxylic group metal oxide framework incorporated (UiO-66-
(COOH)2) ZnIn2S4/MoS2 composite based on Z-scheme photocatalytic mechanism
exposed to be a better approach for heavy metal reduction.
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 607

23.4.4  M
 oS2 as Non-noble Metal-Based Cocatalyst
for Antibacterial Disinfection

Water contamination by microorganisms such as bacteria and viruses can cause


severe health problems. Present conventional techniques such as chlorination,
ozonation, and UV irradiation are not efficient as well as more costly. Therefore, the
development of an efficient and cost-effective method is essential for photocatalytic
antibacterial disinfection. Semiconducting materials such as TiO2 and ZnO have
been employed for photocatalytic disinfection of bacteria, but these are wide
bandgap materials and UV light-active materials. In the solar spectrum, only 5%
consists of UV light, and more than 40% consists of visible light. Therefore,
development of visible light-active materials is more useful. Moreover, MoS2 is
used as the cocatalyst which is loaded with various semiconducting materials for
antibacterial disinfection under visible light irradiation.
The bacterial disinfection process involves bacterial cell wall damage by the bio-
active reactive oxygen and hydroxyl radicals from the redox reaction of charge car-
riers generated by absorbing light energy equal to the bandgap of a material.
In 2020 X. Hu et al. synthesized g-C3N4-loaded MoS2 for bacterial disinfection
under visible light irradiation [31]. g-C3N4 is the most abundant and acts as a typical
semiconductor, but it has some limitations such as a high charge recombination rate
and less charge transfer rate. To avoid this limitation, metallic-phase rich MoS2
loaded as cocatalyst which suppresses the electron-hole recombination and increases
the charge transfer rate causing enhanced production of the reactive oxygen species,
thereby increasing the death rate of bacteria as shown in Fig. 23.10.
In photocatalytic gram-negative bacteria, E. coli is incubated and tested for pho-
tocatalytic disinfection by irradiating 150 W xenon lamp. Three milligrams of the
MoS2-loaded g-C3N4 is added to the 30-mL solution of incubated E. coli bacteria.
For finding the disinfection rate, absorbance is measured for every 20  min of
irradiation. Depending on the concentration, the amount of bacterial death is
analyzed. Within an hour complete disinfection of bacteria was observed for
1T-phase metallic MoS2-loaded g-C3N4. Overall, 91% of bacteria were disinfected
in 40 min, whereas pure g-C3N4 resulted in 84.67% disinfection, which was lower
than 1T-phase metallic MoS2-loaded g-C3N4. The main reason for the efficient
disinfection of bacteria may be due to the enhanced charge transfer rate and a greater
amount of suppression of electron and hole recombination. In 1T-phase MoS2, both
Mo basal plane and outer S atoms are reactive, but in 2H-phase MoS2, only S edges
are reactive MO basal plane totally inactive; therefore, 1T-phase metallic MoS2-­
loaded g-C3N4 provides a greater number of active sites compared to semiconducting
2H-phase MoS2. The conduction band electrons separated by 1T-phase MoS2
cocatalyst tend to reduce more amount of dissolved O2 to produce highly reactive
H2O2. Later, these generated H2O2 participated in disinfection or death of bacteria.
It is very much essential to develop a metallic-phase or conductive MoS2 for
efficient disinfection of bacteria under visible light irradiation, so that we can avoid
highly expensive and nonviable noble metals for bacterial disinfection.
608 M. Muniyappa et al.

Fig. 23.10  The mechanism of ZnIn2S4/MoS2 for bacterial disinfection. Reprinted from [31], with
permission from Elsevier

23.5  Summary

In photocatalytic applications, replacing expensive noble metals as cocatalyst is


very much essential for the sustainable development in the field of energy and
environmental applications. In this chapter, we have discussed basic properties,
synthesis, and various applications of MOS2 as a non-noble metal-based cocatalyst
for photocatalysis. Synthesis strategies like solvo-/hydrothermal and chemical
exfoliation methods were discussed. Recent studies on non-noble metal-based
MoS2-loaded semiconductors of hydrogen generation have studied thoroughly.
More attention was given for strategies for improving the hydrogen generation
process. Other applications of MoS2 as cocatalysts such as photocatalytic dye
degradation, toxic metal ion reduction, bacterial disinfection, and their mechanisms
were discussed.

References

1. Fujishima, A., & Honda, K. (1972). Electrochemical photolysis of water at a semiconductor


electrode. Nature, 238, 38–40. https://doi.org/10.1038/238038a0.
2. Onsuratoom, S., Puangpetch, T., & Chavadej, S. (2011). Comparative investigation of hydro-
gen production over Ag-, Ni-, and Cu-loaded mesoporous-assembled TiO2-ZrO2 mixed
23  Nanostructured MoS2 as Non-noble Metal-Based Cocatalyst… 609

oxide nanocrystal photocatalysts. Chemical Engineering Journal, 173, 667–675. https://doi.


org/10.1016/j.cej.2011.08.016.
3. Radisavljevic, B., Radenovic, A., Brivio, J., Giacometti, V., & Kis, A. (2011). Single-layer MoS2
transistors. Nature Nanotechnology, 6, 147–150. https://doi.org/10.1038/nnano.2010.279.
4. Wang, Q.  H., Kalantar-Zadeh, K., Kis, A., Coleman, J.  N., & Strano, M.  S. (2012).
Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nature
Nanotechnology, 7, 699–712. https://doi.org/10.1038/nnano.2012.193.
5. Wang, X., Maeda, K., Thomas, A., Takanabe, K., Xin, G., Carlsson, J. M., et al. (2009). A
metal-free polymeric photocatalyst for hydrogen production from water under visible light.
Nature Materials, 8, 76–80. https://doi.org/10.1038/nmat2317.
6. Maeda, K., Higashi, M., Lu, D., Abe, R., & Domen, K. (2010). Efficient nonsacrificial water
splitting through two-step photoexcitation by visible light using a modified oxynitride as a
hydrogen evolution photocatalyst. Journal of the American Chemical Society, 132, 5858–5868.
https://doi.org/10.1021/ja1009025.
7. Zhou, W., Yin, Z., Du, Y., Huang, X., Zeng, Z., Fan, Z., et al. (2013). Synthesis of few-layer
MoS2 nanosheet-coated TiO2 nanobelt heterostructures for enhanced photocatalytic activities.
Small, 9, 140–147. https://doi.org/10.1002/smll.201201161.
8. Qin, L.-Z., Lin, Y.-Z., Dou, Y.-C., Yang, Y.-J., Li, K., Li, T., et al. (2020). Toward enhanced
photocatalytic activity of graphite carbon nitride through rational design of noble metal-free
dual cocatalyst. Nanoscale, 12(25), 13829–13837. https://doi.org/10.1039/c9nr10044j.
9. Nagaraja, C.  M., Kaur, M., & Dhingra, S. (2020). Enhanced visible-light-assisted photo-
catalytic hydrogen generation by MoS2/g-C3N4 nanocomposites. International Journal of
Hydrogen Energy, 45, 8497–8506. https://doi.org/10.1016/j.ijhydene.2020.01.042.
10. Yuan, Y. J., Wang, F., Hu, B., Lu, H. W., Yu, Z. T., & Zou, Z. G. (2015). Significant enhance-
ment in photocatalytic hydrogen evolution from water using a MoS2 nanosheet-coated ZnO
heterostructure photocatalyst. Dalton Trans, 44, 10997–11003. https://doi.org/10.1039/
c5dt00906e.
11. Yuan, Y. J., Tu, J. R., Ye, Z. J., Lu, H. W., Ji, Z. G., Hu, B., et al. (2015). Visible-light-driven
hydrogen production from water in a noble-metal-free system catalyzed by zinc porphy-
rin sensitized MoS2/ZnO. Dyes and Pigments, 123, 285–292. https://doi.org/10.1016/j.
dyepig.2015.08.014.
12. Xiang, Q., Yu, J., & Jaroniec, M. (2012). Synergetic effect of MoS2 and graphene as cocata-
lysts for enhanced photocatalytic H2 production activity of TiO2 nanoparticles. Journal of the
American Chemical Society, 134, 6575–6578. https://doi.org/10.1021/ja302846n.
13. Ge, L., Han, C., Xiao, X., & Guo, L. (2013). Synthesis and characterization of composite
visible light active photocatalysts MoS2-g-C3N4 with enhanced hydrogen evolution activ-
ity. International Journal of Hydrogen Energy, 38, 6960–6969. https://doi.org/10.1016/j.
ijhydene.2013.04.006.
14. Chang, K., Mei, Z., Wang, T., Kang, Q., Ouyang, S., & Ye, J. (2014). MoS2/graphene cocata-
lyst for efficient photocatalytic H 2 evolution under visible light irradiation. ACS Nano, 8,
7078–7087. https://doi.org/10.1021/nn5019945.
15. Tian, Q., Zhang, L., Liu, J., Li, N., Ma, Q., Zhou, J., et al. (2015). Synthesis of MoS2/SrZrO3
heterostructures and their photocatalytic H2 evolution under UV irradiation. RSC Advances, 5,
734–739. https://doi.org/10.1039/c4ra11135d.
16. Nos, S., Joensen, P., Frindt, R. F., & Morrison, S. R. (1986). Single-layer MoS2. Mater Res
Bull, 21, 457–461.
17. Frame, F. A., & Osterloh, F. E. (2010). CdSe-MoS2: A quantum size-confined photocatalyst
for hydrogen evolution from water under visible light. Journal of Physical Chemistry C, 114,
10628–10633. https://doi.org/10.1021/jp101308e.
18. Guan, J., Wu, J., Jiang, D., Zhu, X., Guan, R., Lei, X., et al. (2018). Hybridizing MoS2 and
C60 via a van der Waals heterostructure toward synergistically enhanced visible light pho-
tocatalytic hydrogen production activity. International Journal of Hydrogen Energy, 43(18),
8698–8706. https://doi.org/10.1016/j.ijhydene.2018.03.148.
610 M. Muniyappa et al.

19. Zhu, Y., Ling, Q., Liu, Y., Wang, H., & Zhu, Y. (2015). Photocatalytic H2 evolution on MoS2-­
TiO2 catalysts synthesized via mechanochemistry. Physical Chemistry Chemical Physics, 17,
933–940. https://doi.org/10.1039/c4cp04628e.
20. Zhao, Y., Yang, Z., Zhang, Y., Jing, L., Guo, X., Ke, Z., et al. (2007). Cu 2 O decorated with
cocatalyst MoS 2 for solar hydrogen production with enhanced efficiency under visible light.
The Journal of Physical Chemistry, 118, 14238–14245.
21. Nguyen, M., Tran, P.  D., Pramana, S.  S., Lee, R.  L., Batabyal, S.  K., Mathews, N., et  al.
(2013). In situ photo-assisted deposition of MoS2 electrocatalyst onto zinc cadmium sul-
phide nanoparticle surfaces to construct an efficient photocatalyst for hydrogen generation.
Nanoscale, 5, 1479–1482. https://doi.org/10.1039/c2nr34037b.
22. Wei, L., Chen, Y., Lin, Y., Wu, H., Yuan, R., & Li, Z. (2014). MoS2 as non-noble-metal co-
catalyst for photocatalytic hydrogen evolution over hexagonal ZnIn2S4 under visible light
irradiations irradiations. Appl Catal B Environ, 144, 521–527. https://doi.org/10.1016/j.
apcatb.2013.07.064.
23. Zhao, W., Liu, J., Ding, Z., Zhang, J., & Wang, X. (2020). Optimal synthesis of platinum-
free 1D/2D CdS/MoS2 (CM) heterojunctions with improved photocatalytic hydrogen produc-
tion performance. Journal of Alloys and Compounds, 813, 152234. https://doi.org/10.1016/j.
jallcom.2019.152234.
24. Wang, Z., Hou, J., Yang, C., Jiao, S., & Zhu, H. (2014). Three-dimensional MoS2–CdS–β-­-
TaON hollow composites for enhanced visible-light-driven hydrogen evolution. Chemical
Communications, 50, 1731–1734. https://doi.org/10.1039/c3cc48752k.
25. Maitra, U., Gupta, U., De, M., Datta, R., Govindaraj, A., & Rao, C.  N. R. (2013). Highly
effective visible-light-induced H 2 generation by single-layer 1T-MoS 2 and a nanocomposite
of few-layer 2H-MoS 2 with heavily nitrogenated graphene. Angew Chem Int Ed Engl, 52,
13057–13061. https://doi.org/10.1002/anie.201306918.
26. Wei, D., Ding, Y., & Li, Z. (2020). Noble-metal-free Z-Scheme MoS2–CdS/WO3–

MnO2 ­ nanocomposites for photocatalytic overall water splitting under visible light.
International Journal of Hydrogen Energy, 45, 17320–17328. https://doi.org/10.1016/j.
ijhydene.2020.04.160.
27. Shi, X., Fujitsuka, M., Kim, S., & Majima, T. (2018). Faster electron injection and more active
sites for efficient photocatalytic H2 evolution in g-C3N4/MoS2 hybrid. Small, 14, 1–9. https://
doi.org/10.1002/smll.201703277.
28. Kumar, S., Sharma, V., Bhattacharyya, K., & Krishnan, V. (2016). Synergetic effect of MoS2-­
RGO doping to enhance the photocatalytic performance of ZnO nanoparticles. New Journal of
Chemistry, 40, 5185–5197. https://doi.org/10.1039/c5nj03595c.
29. Pudkon, W., Bahruji, H., Miedziak, P.  J., Davies, T.  E., Morgan, D.  J., Pattisson, S., et  al.
(2020). Enhanced visible-light-driven photocatalytic H2 production and Cr(vi) reduction of
a ZnIn2S4/MoS2 heterojunction synthesized by the biomolecule-assisted microwave heating
method. Catalysis Science & Technology, 10, 2838–2854. https://doi.org/10.1039/d0cy00234h.
30. Mu, F., Cai, Q., Hu, H., Wang, J., Wang, Y., Zhou, S., et al. (2020). Construction of 3D hier-
archical microarchitectures of Z-scheme UiO-66-(COOH)2/ZnIn2S4 hybrid decorated with
non-noble MoS2 cocatalyst: A highly efficient photocatalyst for hydrogen evolution and
Cr(VI) reduction. Chemical Engineering Journal, 384, 123352. https://doi.org/10.1016/j.
cej.2019.123352.
31. Hu, X., Zeng, X., Liu, Y., Lu, J., Yuan, S., Yin, Y., et  al. (2020). Nano-layer based 1T-rich
MoS2/g-C3N4 co-catalyst system for enhanced photocatalytic and photoelectrochemi-
cal activity. Applied Catalysis B: Environmental, 268, 118466. https://doi.org/10.1016/j.
apcatb.2019.118466.
Index

A SiO2 substrate, 122


Ab initio molecular dynamics, 342 thermal energy, 121
Ab initio simulations, 339 UV illumination, 122
Absorption capacity, 157 Air pollutants
Activated carbon, 328 diseases, 521
Adsorbent/photocatalyst composite materials, natural/manmade sources, 521
523, 541 photocatalytic destructions (see
Adsorbents, 425 Photocatalytic destructions, air
Adsorption, 330, 332, 340, 344, 353, 384 pollutants)
Adsorption capacity, 332 PM, 522, 523
Adsorption isotherm models, 330 resources, 522
Adsorption isotherms, 347, 357 Air pollution, 521, 522
Adsorption kinetics, 357 Air purification, 45, 53, 540
Adsorption material-integrated All-solid-state Z-scheme heterojunction
photocatalyst, 423 composites
Adsorption processes, 256 Ag-deposited WO3-X nanocapsules, 96
Advanced oxidation process (AOP), 2, BiOCl-Au-CdS heterostructure, 96
149, 329 BiOI nanosheets, 96
Advanced technological developments, 511 charge transfer mechanism, 96
Ag/AgX/GO hybrid nanocomposites, 127 iodine vacancies, 96
AgCl/CNTs/graphitic carbon nitride, 349 photogenerated electron transfer, 95
AgCl surface, 127 RGO, 96, 97
Ag-doped rutile, 118 tetracycline degradation efficiency, 97
Ag doping concentration, 118 3D heterostructure, 97
Ag nanoparticles (AgNPs), 465 2D/2D heterostructure, 96, 97
Agglomeration, 470 ZnFe2O4/Ag/PEDOT materials, 97
Aggregated Ti-oxides, 539 Ambient argon treatment, 272, 273
Ag-loaded nanocomposites, 184 Ambient hydrogen–argon treatment, 271, 272
Ag–metal oxides Ambient hydrogen–nitrogen treatment, 272
Ag NPs, 123 American Society for Testing and Materials
Ag–AgnX, 123 (ASTM), 367
Au/TiO2 a model system, 121 Ammonia, 448
hot electron transfer, 121 Amorphous Cr(III)-Te(IV), 474
L-1 methyl orange solution, 122 Analytes, 565, 571
plasmonic photocatalyst Ag@AgCl, 122 Angular frequency, 171
SB, 121 Anodization method, 275

© Springer Nature Switzerland AG 2021 611


S. Balakumar et al. (eds.), Nanostructured Materials for Environmental
Applications, https://doi.org/10.1007/978-3-030-72076-6
612 Index

Anodization process, 7, 8 AuNPs/B-TiO2/FTO substrates, 466


Anthracene, 310 Aurivillius compounds photocatalytic
Antibacterial disinfection, 607 applications
hydroxyl radicals, 156 antibacterial disinfection, 156–157
photocatalytic disinfection, 156 dye degradation, 148–150
photocatalytic mechanism, 156, 157 heavy metals reduction, 154–156
plasmon-induced, 156 narrow bandgap, 148
radical trapping, 157 organic pollutants degradation, 150–152
reactive oxygen species, 157 pharmaceutical pollutants
semiconducting materials, 156 degradation, 152–154
solar spectrum, 156 Z-scheme heterojunction-based
traditional methods, 156 photocatalysis, 148
Antibiotic ciprofloxacin, 89 Aurivillius-phase structures, 138
Antibiotic enrofloxacin, 94 Aurivillius pristine compounds, 152
Antibiotic groups, 337 Aurivillius structural arrangement, 139
Antibiotic molecules, 87
Antibiotic resistance genes (ARG), 356
Antibiotic resistances, 85 B
Antibiotics Bacterial disinfection, 592, 607, 608
adsorption, 96, 103 Ball milling, 597
chemical removal techniques, 85, 86 Bandgap, 301, 306, 307, 547
chloramphenicol molecules, 103 Bandgap engineering
ciprofloxacin, 88, 89, 105 anti-bacterial agent, 551
degradation, 86, 97, 104, 108, 332 CO, 551
emerging pharmaceutical pollutant, 85, 108 CO2 utilization, 551
levofloxacin removal, 105 efficient degradation, 549
nanocomposite systems, 89 electron reduction, 549
pharmaceutical drugs, 85 Fe3+-modified ZnO, 551
pharmacological properties, 332 FSP, 551
photogenerated electron, 86 IFCT, 550
release, 85 MER, 549, 550
residues, 85 MMCT, 549
sources, 355 ROS, 551
tetracycline, 97 suitable candidate, 550
types, 94 TiO2, 552
Anti-microbial agent, 559 XANES analysis, 549
Antimicrobial applications, 22, 23 Bandgap narrowing, 283
Antimicrobial-resistant organisms, 332 Bandgap vitality, 486
Aquatic environment, 368 Beer-Lambert Law, 408
Arc discharge, 574 Bentonite nanohybrid-modified polyaniline
Argon–nitrogen treatment, 273 (PANI) nanofibers, 578
Artificial photocatalysis, 591 Bi-based Aurivillius compounds
As-grown SWCNTs (As-Fe-SWCNTs), 341 applications (see Aurivillius compounds
Atomic layer deposition (ALD), 495 photocatalytic applications)
Au and CuS nanoparticle-modified TiO2 crystal structure and properties, 158
nanobelt (Au-CuS-TiO2 NBs), 92 layered structures, 138
Au–metal oxides pharmaceuticals degradation, 158
insulating oxides, 123 preparation, 144
interactions, 123 surface modification, 158
photocatalysis, 123 synthesis, 144
photocatalytic hydrogen generation, 124 visible light radiation, 138
UV photochemistry, 123 Bi-based Aurivillius materials, 158
Au-MWCNT/PANI nanocomposites, 189 Bi-based materials, 150
Au nanoparticles (AuNPs), 465 Bi-based molybdates, 142
Index 613

Bi-based photocatalysis materials Black Titania applications


electron-hole pairs, 140 CO2 photo reduction, 286
hydroxyl radicals, 141 hydrogen production, 285
organic contaminants, 140 pollutant removal, 284, 285
oxidation, 141 Black titania surface structure, 283, 284
photocatalytic degradation mechanism, 141 Boron nitride (BN), 488, 506
photoexcitation, 141 Branched-TiO2microrods, 465
prolonged recombination, 140 Bronsted acid sites, 528, 529
Bi-based photocatalysts Buckyballs, 324
surface modification, 146–148
VB, 138
visible light, 138 C
Bimettalic nitride fluorides synthesis C60-based photocatalysts, 349
calcium barium, 499 Cadmium sulfide (CdS), 242, 462
calcium magnesium, 499 Cadmium sulphide-reduced graphene oxide
materials, 499 (CdS-RGO), 462
Biochar, 356 Calcium barium nitride fluoride, 499
BiOCl 3D hierarchical architecture, 88 Calcium magnesium nitride fluoride, 499
BiOCl hierarchical porous Calcium nitride fluoride phase, 497
materials, 88 Calibration curve, 409
Biogeochemical cycles, 355 Capacitive deionization (CDI), 248
Biological removal methods, 338 Carbamazepine (CBZ), 307
Biological sensors, 571 Carbon, 322, 433
Biomagnification, 355 Carbonaceous nanomaterials
Biosensing hydrogen generation, 459 advantages, 329
Biosensors carbon, 322
electrochemical sensor, 583 CNTs, 324, 357
electrodes, 583 environmental applications,
environmental monitoring, 583 322, 356
sensing mechanism, 572 fullerenes, 324
ultrasensitive electrochemical functionalization, 337
sensor, 583 gas pollutants removal, 353–354
Bisphenol, 344 graphene, 323
Black TiO2 nanomaterials, 286 membranes, 350
Black TiO2 preparation methods physico-chemical properties, 328
chemical reduction, 275–276 properties, 324
electrochemical synthesis, 276 soil remediation, 355–357
imidazole reduction, 278 sulfamethoxazole, 335
ionothermal synthesis, 277 synthesis, 325–329
plasma treatment, 274 synthesized, 323
proton implantation, 278 toxicity, 357
pulsed laser ablation, 277 Carbon allotropes, 323, 433
Si chemical etching method, 278 Carbon atoms, 322, 574
thermal treatment, 269–274 Carbon-based materials, 228, 356
Black TiO2 properties Carbon-based nanodots, 324
HRTEM, 279, 280 Carbon-based nanomaterials, 357, 567, 568
preparation approaches, 283 Carbon-based nanostructure, 580, 581
SAED, 279 Carbon bonding composition
semiconductor metal oxide, 278 ratios, 226
surface disorders, 279 Carbon dots, 324
surface functional groups, 281 Carbon monoxide (CO), 521
surface modifications, 282 Carbon nanohorn (CNH), 324, 326
Ti3+ and oxygen vacancies, 280 Carbon nanomaterials, 330, 355, 356
trapping centers, 283 Carbon nano-onions, 324
614 Index

Carbon nanotubes (CNT), 311, 572 Ciprofloxacin, 332–334


CNH, 326 Citric acid, 9
common terminal group, 326 Clean drinking water, 312
functionalization, 326 Climate change, 455
morphologies, 326 CMC-Fe(II) complex, 251
production techniques, 326 CNT delivers, 352
types, 326 Cobalt nitride fluoride system, 498
Carbon nitride (CN), 501 Cocatalysts
Carbon nitride polymer (CNP), 491 Cds, 599
Carbon quantum dots (CQDs), 102, 104, conductivity, 603
327, 433 metallic MoS2, 601
Carbon sphere-based composites MoS2, 597–599
direct dual semiconductor noble metals, 592
photocatalyst, 105 photocatalytic applications, 608
hydrothermal reaction, 105 physical properties, 603
N-HMCs, 106 semiconducting metal oxides, 592
photocatalytic material designing, 105 Colorimetric method, 410
spheres/g-C3N4 composites, 105 Combustible hydrogen fuel production, 114
Carbon-rich C3N4 photocatalyst, 491 Commercial RP (C-RP), 458, 459
Carboxymethyl cellulose (CMC), 251 Common modified Hummers method, 220
Carcinogenic organic compounds, 150 Conducting polymers, 569
Catalytic H2 generation, 511 Conduction band (CB), 128, 283, 300, 311,
CBZ photodegradation, 308 458, 546, 597
C-doped ZnO photocatalyst, 306 edge potentials, 593, 599
CdS–graphene composites, 462 electrons, 607
CdS-RGO composite, 463 Conduction band minimum (CBM), 118
Ce-metal-organic framework-808, 73 Conductive carbonaceous materials, 600
Cetyl trimethyl ammonium bromide (CTAB), Conjugate polymer-based sensors, 578
6, 69, 70, 147, 302 Connection coefficient (R2) estimation, 492
Charge carrier recombination, 532 Contaminated sediments, 259
Charge carrier separation, 434 Contaminated sewage water, 152
Charge separation efficiency, 230 Continuous flow system, 425
Charge separation mechanism, 227 Controlled nucleation, 217
Chemical bath deposition (CBD), 495 Conventional and emerging pollutants, 329
Chemical energy conversion, 457 Conventional cellulose acetate-based
Chemical exfoliation, 595, 596, 608 membranes, 350
Chemical functionalization, 325, 332, 351 Conventional clean-up techniques, 247
Chemical nanosensors, 566 Conventional treatment methods, 332
Chemical oxygen demand (COD), 409, 410 Conventional UV–vis spectroscopy
Chemical precipitation, 144 monitoring, 314
Chemical reactions, 5 COOH group gatherings, 493–494
Chemical sorption method, 338 Copolymerization, 493
Chemical stability, 217, 227 Core/shell mesoporous hybrid-structured
Chemical treatments, 346 TiN-VN nanofibers, 500
Chemical vapor deposition (CVD), 325, 436, Core/shell/shell γ-Fe2O3/microporous SiO2/
495, 572, 573 polypyrrole microspheres, 204
Chironomus riparius, 422 Coulomb attraction, 116
Chlorotetracycline, 333, 337 Coupling graphene, 445
Chromate infectedwastes, 473 Covalent parallel nitrides, 488
Chromatereducing microscopic Cr containing industrial wastewater, 473
organisms, 473 Cr(III) metals
Chromium (Cr) accumulation, 474
metal alloys, 456 human use, 456
oxidation forms, 456 immobilization, 475, 476
Index 615

roles, 456 LaFeO3 nanoparticles, 99


utilizing planktonic cells, 475–476 mediator materials, 98
Cr(VI) metals tetracycline degradation efficiency, 99
application areas, 456 2D g-C3N4 nanosheets, 100
bioreduction, 474 WO3-g-C3N4 heterojunction composite, 98
bioremediation, 475 Direct/indirect oxidation, 141
chemical reduction, 457 Dominant shielding mechanism, 201
concentration, 457 Doping, 327, 555
detoxification, 457 Dosage, 391
reduction, 460 Dual functional adsorbent/photocatalyst
removal from water, 457 system, 540
toxic reduction, 477 Dubinin–Radushkevich models, 339
water contamination, 457 Dye adsorption, 401
Cr(VI) reduction reduction, 458 Dye degradation, 2, 5, 6, 9, 11, 12, 33, 38–42,
Cr(VI) untreated cells, 476 44, 45, 49, 53, 74, 75
Cryogenic distillation, 353 advanced oxidation process, 149
CTAB-modified carbon spheres, 105 charge recombination, 149
CTCN heterojunction, 310 charge separation process, 149
c-type cytochromes, 474 heterostructures, 150
Cu2+ ion-modified Ti3+ self-doped TiO2 molecules, 151
material, 554 radicals, 149
Cyanamide/dicyandiamide polycondense, 493 2D Z-scheme-based catalysts, 149
Cyclic voltammetry technique, 500, 583 water pollution, 148
Z-scheme-type photocatalysts, 149
Dye-contaminated water, 412
D Dyes, 340
DBP molecule, 306 Dye sensitization, 290
Decarboxylation, 307 Dye-sensitized solar cell (DSSC), 495
Dechlorination activity, 309
Decomposition, 522, 525, 527
Defect-free crystallite structures, 511 E
Defluoridation, 384, 391 EDA interaction, 334
Degradation efficiency, 354 EDC degradation, 313–314
Dehalorespiration, 257 EDC mineralization, 312
Density functional theory (DFT), 117, 139 EDC passing, 314
Density of states (DOS), 284 EDC photodegradation, 309
DEP degradation pathways, 305 EDTA light-driven photocatalyst, 147
Desalination, 248, 350 EELS spectroscopy, 225
DFT calculations, 118 Electrochemical oxidation, 7
DFT studies, 348 Electrochemical photolysis, 403
Dibutyl phthalate (DBP), 306 Electrochemical processes, 256
Dihydroxybenzene, 307 Electrochemical sensor, 578
Dimethyl phthalate (DMP), 302 Electrochemical synthesis, 276
Dimethylformamide (DMF), 504 Electromagnetic (EM) wave
Direct oxidation, 51 absorption, 169
Direct Z-scheme heterojunction composites electric field, 167
advantages, 98 frequency radiation, 167
cephalexin degradation, 100 incident, 169
charge migration, 98 measuring, 169
CuInS2/Bi2WO6 heterojunction, 99 reflection, 169
degradation experiments, 100 shielding material, 170
electron-hole pair recombination, 99 Electromagnetic absorption, 204
g-C3N4 nanorod, 99 Electromagnetic interference (EMI), 167
hydrothermal method, 100 Electromagnetic radiation, 71–72
616 Index

Electromagnetic spectrum, 113 Endocrine system, 299, 300


Electromigration, 256 Energy fuels, 77
Electron, 548 Energy-level matching, 555–558
Electron charge transfers, 529 Energy renovation techniques, 137
Electron donor acceptor (EDA), 330 Enhanced CO2 adsorption, 442
Electron exchange procedure, 503 Enhanced H2 generation, 505
Electron microscopy, 474 Enhanced photocatalytic activity, 448
Electron spin resonance (ESR), 153, 280 Enhanced photostability, 443
Electron trapping, 43 Enhanced plasmonic photoelectrochemical
Electron-hole recombination, 144, 152 activity, 124
Electron-hole separation barrier, 156 Environmental applications, 247
Electronic pathways, 449 Environmental contamination, 567
Electrons, 2 Environmental decontamination, 440
Electrospinning, 8, 9 Environmental monitoring, 565, 567, 575,
Electrospray ionization mass spectrometry 578, 582–584
(ESI-MS), 306 Environmental pollution, 126, 321, 528
Electrostatic interactions, 338, 347 Environmental remediation, 1–3, 33–35, 40,
Emerging contaminants, 329 53, 137, 321
EMI shielding Environmental sensors, 565, 566
absorption, 171 ESR signals, 39, 41
candidates, 205 Ethyl paraben (EP), 310
core@shell materials, 201–205 Ethylene glycol (EG), 73, 144
efficiency, 172 Ethylenediaminetetraacetic acid (EDTA), 147
electrical conductivity, 171 Exchange-correlation function, 117
ICP-coated fabrics (see PPY-coated fabrics) Exfoliation, 435
ICPs, 172
industrial needs, 167
metallic materials, 167 F
multiple reflections, 171 Fabricated sensor, 581, 582
nanocarbon-based materials, 187–193 F-Ag-β-CD/TiO2/AC photocatalyst, 311
reflection loss, 171 Fe(II)-catalyzed processes, 252
Endocrine disrupting chemicals (EDC) Fe3+ ion-induced IFCT, 551
binary nanostructures, 306–310 Fe3O4/MWCNT/wax composites, 190
concentration, 299 Fe-Cr hydroxide, 254
CTAB, 302 Fe-doped vacancy-defected CNTs, 354
decomposition, 300, 314 FeNi nanoparticles, 186
harmonically active environmental Fenton’s process, 2
pollutants, 300 Fenuron, 339
homogeneous/heterogeneous Ferrite-based magnetic nanoparticles, 342
photocatalysis, 300 Ferrofluid-based nanoarchitectured PPF, 196
hormones travel, 299 First-order kinetic expression, 408
photocatalytic activity, 301 Flame spray pyrolysis (FSP), 551
photocatalytic degradation, 303, 305 Fluoride, 383
photocatalytic performance, 306 Fluoride adsorption studies, 388
physical treatment, 300 dosage, 391
single component photocatalysts, 302 equilibrium time, 392
target substrates, 300 fluoride concentration, 389
ubiquitous, 300 fluoride ion, 389, 393
ZnO, 303 FTIR spectra, 393
Endocrine disruptors IEP and pH, 394
decomposition, 302 ion-exchange mechanism, 393
molecular structures, 301 ion-selective electrode, 388
World Wildlife Fund Wingspread pH, 392
Conference in 1991, 300 protonic trititanate nanostructures, 393
Index 617

surface area, 390 Geometrics-dependent optical and electronic


TNTs, 390 properties, 4
zeta potential, 392 GH-CIP-Cu(II) ternary complex, 334
Fluoride removal, 33 Global environmental issues, 321
Fluoroquinolones, 332 GO aqueous suspension, 221
Fly ash cenosphere (FAC), 185 GO composite, 342
Fourier transform infrared spectroscopy GO contains, 330
(FTIR), 393 GO hydrophilicity, 330
Free space, 171 GO membrane, 352
Freundlich isotherm, 337, 342 GO sulfamethoxazole (GO-SMZ), 356
Friedel-Crafts response, 490 GO/TiO2 composite, 337
Fullerenes, 324, 327, 348 GR allotrope, 438
Functionalization, 328 Graphene (GR), 323, 332, 351
Functionalization possibilities, 326 carbon allotropes, 433
Functionalized CNTs, 341, 346, 354 modified forms, 433
Functionalized magnetic fullerene properties, 437, 438
nanocomposites, 343 SB, 434
Functionalized SWCNTs, 344 surface electrocatalytic reactions, 434
Furnace temperature, 573 synthesis (see GR synthesis)
Graphene-based composites, 118
Graphene-based nanoadsorbents, 330
G Graphene-based nanomaterials, 353
Gadolinium nitrate (GdNO3), 9 Graphene-based photocatalytic materials,
Gallium nitride (GaN), 506 449, 450
GaN/Si nanowires, 507 Graphene hydrogels (GH), 334
Gas chromatography mass spectrometry, 302 Graphene loading, 193
Gas sensor Graphene oxide (GO), 222, 326, 356, 438
applications, 578, 584 Graphene oxide core/shell structures, 224
CO, 576 Graphene oxide/metal oxide (GO/MO)
H2 gas sensor, 580 heterojunction, 47–49
and humidity-detecting sensors, 569 Graphene oxide-modified lanthanum vanadate
nanomaterial-based toxic, 566 (GO/LaVO4), 103
NH3 gas sensor, 570 Graphene quantum dots, 324, 348
PANI nanofibers, 578 Graphene-MnFe2O4 nanocomposite, 339
sensing mechanism, 570 Graphitic carbon nitride (g-C3N4), 128, 510–511
silicon-based, 569 carbon nitrides, 489
sol-gel method, 579 hydrothermal stability, 489
toxic gases (see Toxic gas sensor) nanosheets, 45
Gaseous pollutants, 357 heterojunction, 46, 47
Gas-sensing property, 580 hybrid photocatalysts, 493, 494
Gas-sensing reaction, 576 GR-based photocatalysts applications
Gas-sensing system, 580, 581 CO2 photoreduction, 442, 444
g-C3N4/Ag hybrid photocatalysts, 504 heavy metal ions removal, 447, 448
g-C3N4-based hybrid photocatalysts, 508, 509 hydrogen production, 441, 442
g-C3N4-based photocatalysts, 494 nitrogen fixation, 448, 449
g-C3N4-based plasmonic photocatalysis NOx removal, 444, 445
Ag/g-C3N4 nanosheets, 128 organic compounds photodegradation,
MO, 128 440, 441
photocatalytic applications, 128 photocatalytic disinfection, 445, 446
photocatalytic efficiency, 128 water splitting, 441, 442
VB, 128 GR-based photocatalytic nanocomposites, 438
GC-MS measurement, 303 GR-based semiconductor photocatalysts
GC-MS technique, 302 preparation, 438
g-CN/G/RCF composites, 509 GR-based systems, 441
618 Index

Greenhouse gases, 77 Hierarchical semiconducting


Ground-level ozone (GLO), 521 photocatalyst, 115
Hierarchical surface structure
photocatalyst, 303
H High-efficiency particulate air (HEPA), 545
Harmless products, 524, 541 High energy consumption, 18
Harsh reaction conditions, 18 High gradient magnetic separator
H-doping, 283 (HGMS), 370
Heat, ventilation and air-conditioning High-performance liquid chromatography
(HVAC), 545 (HPLC), 405–408
Heavy metal ions removal, 447, 448 High-performance volatile ethanol
Heavy metal reduction, 19–21, 591, 605, 606 sensors, 577
Heavy metals High-pressure hydrogen treatment, 269
Cr (VI) removal, 423, 424 High-resolution transmission electron
hydroxyl radicals, 423 microscopy (HRTEM), 279
industrial wastewater, 423 High surface area, 566, 579
oxidation/reduction, 423 High temperature, 574
photocatalysis, 423 High-temperature calcination, 554
reduction process, 424 Hollow boron nitride spheres (HBNS), 505
TiO2, 423 Hollow Fe2O3-TiO2-PtOx heterojunction
traditional wastewater treatment, 423 photocatalyst, 241
types, 455 Hollow polypyrrole (HPPy), 184
Heavy metals reduction, 154–156 Hollow TiO2 microspheres (HTMs), 222, 228
Hemispherical nanorods, 490 Horizontal furnace, 573
Herbicides, 339 Household cooking, 545
Heteroatom doping, 150, 327 HPLC measurements, 406
Heterogeneous photocatalysis, 301, 348 HR-TEM images, 225
Heterojunction, 146, 599, 600 Human health effects, 167
Heterojunction photocatalysts Human productivity, 545
2D-0D (see 2D-0D heterojunction Humic acid (HA), 335
photocatalysts) Humidity sensor, 582, 583
types, 34 Hybrid carbonaceous nanomaterials, 343
Heterojunction semiconductors, 599 Hybrid photocatalysts, 425
Heterojunctions, 301 Hybrid-structured metal nitrides applications
photocatalysis (PC), 33 hydrogen generation, 502
photoelectrocatalysis (PEC), 33 photocatalytic degradation
Heterostructure formation, 424 mechanism, 503–505
Hexagonal boron nitride (h-BN), 506 photocatalytic water splitting, 502, 503
Hexavalent chromium, 456 Hydro-/solvothermal reactions, 236
Hierarchical nanostructure Hydro-/solvo-thermal synthesis, 5, 6
applications, 66 (see also Photocatalytic Hydrocarbons dehalogenation, 253
applications) Hydrogen bonding, 330, 336, 344
bulk and surface properties, 65 Hydrogen evolution, 229
characteristics, 67 Hydrogen evolution reaction (HER), 495
compositional variation, 66 Hydrogen peroxide (H2O2), 153
metal oxide semiconducting materials, 65 Hydrogen reduction-based approach, 283
metal oxide semiconductors, 66 Hydrogenation-based treatment, 286
nanometre regime, 67 Hydrogen-doped black titania, 119
photocatalytic performance, 67 Hydrophobic GO-based membrane, 352
physical integration/chemical bonding, 66 Hydrophobic interactions, 338
physicochemical properties, 65 Hydrophobic modification, 352
preparation strategies (see Metal-organic Hydrothermal method, 69, 70, 463
framework (MOF) preparation Hydrothermal synthesis process, 5
strategies) Hydrothermal/solvothermal method,
structural modification, 66 142–144, 594
Index 619

Hydrothermal/solvothermal reaction, 236 capabilities, 248


Hydroxyl groups, 343, 470 core-shell structure, 248
Hydroxyl radicals, 141 environmental treatment field, 248
heterogeneous Fenton-like processes, 249
nanoparticles synthesis (see Nano-zero-­
I valent iron particles (nZVI))
Ibuprofen (IBP), 153, 421 soil pollution, 248
ICP-based functional core-shell Iron oxide nanostructures, 8
composites, 201 Iron oxide precipitation, 256
ICPs nanocomposites Iron oxides, 169
composite conductivity, 196 Isoelectric point (IEP), 389, 392, 394
dielectric materials, 196 IVB-VIB nitrides, 495
Fe3O4 nanoparticles, 196
heterostructures, 196
microwave absorbing performance, 196 L
microwave absorption, 196 Langmuir adsorption isotherm model, 334
multi-layered graphene, 197 Langmuir–Hinshelwood equation, 354
PANI/BaTiO3 composites, 196 Langmuir model, 333
synergetic improvement, 197 Laser ablation, 573–575
Imidazole reduction, 278 L-cysteine, 251
Immobilization, 473, 474 LDM membrane, 351, 353
Impedance mismatch, 171 Levofloxacin antibiotic complex, 153
In situ generated hydrogen, 278 Lewis acid sites, 528, 529
In situ polymerization/hydrothermal Ligand to metal charge transfer process
reaction, 193 (LMCT), 234
In situ remediation strategies, 258 Light-induced degradation, 140
In situ synthesis process, 555 Liquid-phase synthesis, 67
Indigo carmine (IC), 410 Lithium (Li), 487
Indoor environments, 545 Local SPR phenomenon, 491
Industrial waste, 426 Localized surface plasmon resonance
Industrialization, 329, 455 (LSPR), 43
Infrared (IR) region, 367 carrier concentration, 120
Infrared radiation, 113 colloidal metals, 116
Ink-jet printing technique, 173 distinctive technique, 115
Inorganic mesoporous oxide, 328 effects, 115, 120
Inorganic metals, 346 hydrogen-doped ZnO, 120
Integrated circuit (IC) techniques, 569 impacts, 465
Interfacial charge transfer (IFCT), 550 metallic nanoparticles, 116
Interfacies, 256 noble metals, 118
Interstitial alloys, 495 oscillation frequency, 116
Interstitial nitrides, 487 photocatalytic field, 116
Intraparticle diffusion, 334 SP, 116
Intrinsically conducting polymers (ICPs) spherical metallic nanoparticle, 115
conducting materials, 169 TiO2, 120
conventional EMI shielding Low conductivity, 598
materials, 167 Low-density polyethylene (LDPE), 15
EMI shielding applications, 168 Lower concentration molecules, 566
magnetic dielectric materials, 168 Low-intensity indoor lighting, 547
natural/synthetic fibres textiles, 169 Low-pressure hydrogen treatment, 270
oxidative chemical polymerization, 168
types, 167
Ionic nitrides, 487 M
Ionothermal synthesis, 277 Magnesiothermic reduction, 285
Iron nanoparticles application, 257–259 Magnesium nitride fluoride system, 498
Iron nanotechnology Magnetic photocatalysis process, 370
620 Index

Magnetic photocatalysts, 371 maximum photocurrent value, 507


Magnetic recycling nanobelt structure, 505
electromagnetic separation, 377 Ni nanoparticles, 506
humic acid, 377 TiOxNy nanoparticle, 506
magnetic material, 379 Metal nitride fluoride
photocatalysts, 377 pseudo-oxides, 495
photocatalytic degradation, 377 single, 496, 497
separation, 377 thermal stability, 498
suspension solution, 377 TMNs characteristics, 497
Magnetite, 253 Metal nitride photograph impetuses, 490
Manganese-loaded Prussian blue analogs Metal nitrides, 486
(Mn-PBA), 333 Metal organic frameworks (MOF), 65,
Material synthesis, 114 333, 495
Maximum contaminant level (MCL), 383 advantages, 234
Mayan civilization, 116 catalysis applications, 234
Membrane-based technologies, 350 characteristics, 67
Membrane distillation, 352 fabrication, 234
Membrane engineering, 351 hierarchical nanostructures, 234
Membranes, 350 hierarchical structures, 242
Mesoporous g-C3N4, 493 hydrothermal method, 69, 70
Mesoporous materials microwave treatment, 71–73
catalytic and photocatalytic photocatalytic performances, 67
applications, 534 porous materials, 233
Cr-HMS, 537 precipitation method, 68, 69
high surface area, 534 rational fabrication, 234, 235
NO photocatalytic destruction, 535, semiconductor photocatalysis
537, 538 solvothermal method, 70, 71
photocatalytic destruction, 536 synthesis approach, 73, 74 (see also MOFs
selective photocatalysts, 534 visible light semiconductor
solar energy resource, 536 photocatalysis)
tetrahedrally coordinated Ti-oxide synthesis methods, 236, 237
species, 535 2D structure, 242
Ti metal contents, 535 visible light semiconductor
Ti-MCM-41 and Ti-HMS, 534, 535 photocatalysis, 233
visible light irradiation, 538 Metal oxidation/reduction, 423
V-Ti-MCM-41 catalyst, 536 Metal oxides, 65, 66, 68, 71, 72, 74, 81
Metal carbides, 90 nanostructures, 575
Metal doping, 267 sensing applications, 578
Metal hydroxide matrix solution, 5 sensor application, 568
Metal-incorporated ICPs sensors, 569
alloys, 186 solid nanoparticles, 568
Fe-, Co- and Cu, 184, 186 Metal oxide-based nanomaterials, 579, 580
Metal ions, 73, 256 Metal oxide-based semiconductors, 152
Metal nitride-based hybrid photocatalysts Metal oxide-based sensors, 575
BN, 506 Metal oxide semiconductors, 66, 465–468
catalytic activity, 506 Metal oxide-graphene, 468–470
electrocatalytic performance, 505 Metal oxide-metal oxide (1D-0D)
GaN, 506 heterojunction photocatalysts
GaN/Si nanowires, 507 environmental remediation, 40
g-C3N4-based, 508, 509 insulators/metals, 36
H2 production, 505 organic dye degradation applications, 40
h-BN nanosheets, 506 semiconducting metal oxides, 36
HBNS, 505 TiO2-Ag, 41, 42
ITO/V-doped TiO2 sample, 507 TiO2-Ag2O, 38, 40
Index 621

TiO2-Au, 42, 43 electron-hole pair separation, 237


TiO2-Cu, 44, 45 hierarchical structure, 237
TiO2-CuO, 38, 39 hierarchical structured metal oxides,
TiO2-MnO2, 37 240, 241
TiO2-Pt, 43, 44 hierarchical structured metal sulfides, 242
TiO2-ZnO, 39–41 hybrid photocatalysts, 237
Metal oxides semiconductors photocatalysis morphologies, 240
charge carrier separation, 230 photocatalytic CO2 reduction, 239
energy conversion, 223 rGO introduction, 237
graphene, 218 Ru-MOF, 239
rGO matrix, 223 Moisture sensitive, 497
wrapping approach, 230 Molecular dynamics simulation, 352
Metal oxynitrides/sulphides, 65 Molecular dynamics studies, 342
Metal reduction method, 275 Molecular organic frameworks (MOFs)
Metal slats, 410 BT, 472
Metal sulfides, 233, 242 coordination polymers, 470
Metal sulphide semiconductors, 459–462 pillared-layer method, 472
Metal-to-metal charge transfer (MMCT), 549 potential photocatalyst, 472
Methanol oxidation reaction (MOR), 495 RhB organic dye, 471
Methyl orange (MO), 127, 128, 225, 341 Molecular oxygen, 547
Methylene blue (MB), 147, 225, 412, 556 Molecular simulation studies, 354
Methylene blue degradation, 144 Molybdenum disulfide (MoS2)
Micro-iron particles, 250 applications, 592
Micron-sized adsorbents, 384 crystal structure and basic
Micropores, 328 properties, 593
Microporous zeolites electrical properties, 593
structure and composition, 528 monolayer band structure, 594
zeolite-based photocatalysts, 528–531 monolayer MoS2, 593 (see also
Micro-sized sensing particles, 566 Nanostructured MoS2)
Microwave absorption mechanism, 204 as MX2, 593
Microwave heating conditions, 145 synthesis (see Noble metal-free MoS2-­
Microwave plasma-enhanced chemical vapor based nanocomposites)
deposition (MPECVD), 572 three-dimensional structure, 593
Microwave treatment, 71–73 Monobenzyl phthalate, 303
Microwave-transparent interweave Monobutyl phthalate, 303
spacing, 168 Monocrotophos (MCP), 307
Minimally abandoned atomic orbital Morphological properties, 7
(LUMO), 501 Morphology, 3–5, 8–10, 15
Mixed metal oxides synthesis, 217 Morphology-dependent photocatalytic
M-O-C bond, 230 activity, 240
Modified Hummers method, 221 Multi-air pollutants, 522
Modified mesoporous carbon (CMK-3), Multi-component nanomaterials, 301
345, 346 Multi-electron reduction (MER), 549,
Modified polyacrylonitrile substrates, 352 550, 556
MOF-derived hierarcical structured metal Multifunctional hybrid-structured metal
oxides, 240, 241 nitrides
MOF-derived hierarchical structured metal CN, 501
sulfides, 242 TiN-VN, 500
MOF-directed synthesis approach, 73, 74 Multilayer graphene, 351
MOFs visible light semiconductor Multilayer ITO/V-doped TiO2 sample, 507
photocatalysis Multilayer structured cuprammonium fabric/
Ag nanocubes, 240 polypyrrole/copper (CF/PPy/
charge carrier separation, 237 Cu), 184
electron acceptors, 237 Multipulse anodization method, 276
622 Index

Multiwalled carbon nanotubes (MWCNTs), Nanorods, 34


169, 324, 437, 572–574, 582 Nanoscience and nanotechnology
Mutagenic behaviour, 455 advancements, 129
MWCNTs/GO composite, 582 Nanosensors
carbon-based, 571
classification, 566
N environmental application, 566
N-and S-doped C-encapsulated CeO2 Nanosheets
hinge-like hierarchical 2D nanosheet-metal (2D-0D) (see 2D
nanostructures, 73 nanosheet-metal (2D-0D)
Nanoadsorbents, 330 heterojunction photocatalysts)
Nanobelts, 34 Nanosized biochar, 340
Nanocomposite materials, 88 Nano-sized carbon powder, 489
Nanocomposite photocatalyst systems, 86 Nano-sized metal nitrides, 511
Nanocomposite photocatalysts Nanosized red phosphorus, 458
materials, 88, 89 Nanostructure fabrication, 87
semiconductor-carbon, 102–108 Nanostructured hybrid materials, 217
semiconductor-metal composite, 89–92 Nanostructured metal oxides, 579
semiconductor-semiconductor (see Nanostructured MoS2
Semiconductor-semiconductor hydrogen generation
composites) heterojunction nanocomposites, 599, 600
Nanocomposite semiconductor, 301 morphology control, 603
Nanocomposite semiconductor phase engineering, 600, 601
photocatalysts, 87 Z-scheme-based MoS2 nanocomposites,
Nanocomposites, 568, 569, 576, 578, 582, 583 601, 602
Nanocrystals, 155 as non-noble metal-based cocatalyst
Nanocube/graphene oxide (CeM/GO), 103 for antibacterial disinfection, 607
Nanodiamond, 323, 328, 568, 582 for dye degradation, 603–605
Nanofiltration techniques, 352 for hydrogen generation, 597
Nanoflowers, 459 for reduction of toxic metals, 605, 606
Nanomaterial-based toxic gas sensors, 566 Nanostructured photocatalyst, 87–89
Nanomaterials, 247, 258, 329, 349, 426, 566 Nanostructured TiO2 photocatalysts, 523
adsorption, 386 Nanostructures, 2–4
alumina, 386 material properties, 34, 35
carbon-based, 567, 568 Nanotechnology, 247, 321, 385
design and fabrication, 386 Nanotechnology-based solutions, 247
evolution, 567 Nanotubes, 34, 37, 42, 44, 53
metal and metal oxide, 568, 569 Nanowires, 576
nanoscience, 386 Nanoyarns, 459
nanostructure, 567 Nano-zero-valent iron particles (nZVI)
polymer-based, 569, 570 agglomeration, 255
properties, 567 approaches, 249
silicon-based, 569 batch-synthesis, 250
single/hybrid/combination bottom-up pathway, 250, 252
nanostructures, 575 characteristics, 253–256
surface properties, 386 CMC, 251
TiO2-based nanostructures, 386 materials/reagents, 250
TiO6 octahedral units, 386 metallic iron, 251
Nanomembranes, 353 nanosized metallic iron particles, 252
Nanoparticle-based coloring agent, 116 natural extracts, 252, 253
Nanoparticles agglomeration, 255 physical method, 249
Nanoporous carbon materials, 331 sodium borohydride, 251
Nanoporous graphene membrane, 351 soil remediation, 261
Nanoremediation, 258 synthesis, 251
Index 623

National Nanotechnology Initiative, 247 Non-noble transition metals, 233


Natural graphite flakes (NGF), 189 Non-photocatalyst
Natural polymers, 569 photocatalysts, 376
Natural solar light, 75 photocatalytic material, 376
Natural substances, 475 UV light irradiation, 376
Natural sunlight, 412 Norfloxacin, 94
Nb doping, 555 Notable antibiotics, 355
n-butyl benzyl phthalate (BBP), 302 Noticeable light, 491
N-doped Bi2O2CO3/graphene quantum Novel Quinone/Graphene Oxide
dot, 445 Composite, 474
N-doped graphen@polyaniline @ NOx photocatalytic decomposition, 528,
Fe3O4nanocluster, 203 530, 531
N-doped MWCNT, 335 air/semiconductor interface, 531
Near-field enhancement property, 120 antibonding orbitals, 531
Neutron diffraction techniques, 496 composite materials, 528
NiO co-catalyst, 502 impregnated Ti/zeolites, 530
Nitrides photocatalytic abatement, 531
adsorption kinetics, 492 photocatalytic activity, 530
BN, 488 photocatalytic performance, 530
covalent, 488 photodegradation oxidation, 532, 533
interstitial, 487 reaction pathways, 532
ionic, 487 silica/zeolites, 530
nitrogen-containing compound, 486 Y zeolite, 530
principal methods, 487 NOx/VOCs photocatalytic destruction, 540
sulfur, 488, 489 NPs semiconductor nature, 462
surface reinforcing process, 487 NQ-GO material, 476
Nitrogen, 496 Nuclear magnetic resonance (NMR), 281
Nitrogen-doped graphene quantum dots nZVI challenges
(NGQDs), 94, 103 environmental chemistry/
Nitrogen-doped hollow mesoporous carbon geochemistry, 260
spheres (N-HMCs), 106 environmental impact and safety, 260
Nitrogen (N2) fixation, 18–20, 448, 449 production and storage, 260
Nitrogen oxides (NOx), 521, 523 quality control/assurance
N-N-dimethylformamide, 277 standardization, 259
NO photocatalytic destruction, 539, 540 synthesis, 259
NO photocatalytic reaction, 539 nZVI core, 249
NO2 generation, 444 nZVI marine sediment decontamination, 259
Noble metal-free MoS2-based nanocomposites nZVI shelf-life, 258
ball milling, 597 nZVI shell, 249
chemical exfoliation, 595, 596 nZVI treatment, 259
hydrothermal/solvothermal
method, 594–596
Noble metal-free Ti-MOF, 237 O
Noble metals, 579, 592 Occupational safety and health administration
Non-dye-based organic compounds, 150 (OSHA), 456
Non-ionic dyes, 340 One-dimensional structures, 569
Non-magnetic OMWCNT, 339 One-pot hydrothermal method, 576, 579
Non-metal doping, 268, 524 Optical absorption, 119
Nonmetal semiconductor-red phosphorous, 458 Optical bandgap, 592
Non-metallic elements, 548 Optical properties, 592
Non-noble metal Optical sensors, 566
MoS2 (see Molybdenum disulfide (MoS2)) Optical transition mechanisms, 119
Non-noble metal-based MoS2-loaded Ordered mesoporous carbon (OMC), 328, 333
semiconductors, 592, 608 Organic and inorganic substances, 299
624 Index

Organic compounds adsorption, 346 Particulate matter (PM), 521, 523


Organic compounds photodegradation, Pathway management, 258
440, 441 Pd/GCN photocatalysts, 309
Organic contaminants, 33, 148 Pechini method, 9
Organic dye degradation, 33 Pentachlorophenol, 339
Organic dye pollutants, 411, 412 Perfluorooctanesulfonate (PFOS), 343
Organic pollutants, 248, 253, 330, 357 Perovskite-based photocatalysts, 65
aquatic environment, 368 Pesticides, 338
EDCs, 368 chemical compound, 415
PFAS, 368 degradation, 416
POPs, 368 endocrine disrupting activity, 417
WWTPs, 369 fenarimol, 417
Organic pollutants degradation, 331 GO, 416
Organic soil contaminants, 355 health effects, 416
Organic substances, 539 immobilization technique, 416, 417
Organophosphorus herbicide, 338 insecticides, 415
Oxidation-reduction reactions, 113 methyl parathion, 416
Oxidation state, 302 real-life wastewater, 417
Oxidative stress, 357 TiO2 photocatalyst, 416
Oxide photocatalysts, 502 toxic, 415
Oxidized MWCNT (OMWCNT), 339 Pesticides degradation, 338–340
Oxo-bridged redox centre, 549 PET fabric/PPy composites, 177
Oxygen evolution reaction (OER), 495 Pharmaceutical metabolite, 419
Oxygen reduction reactions, 558 Pharmaceutical pollutant degradation, 12–14
Oxygen vacancies Pharmaceutical pollutants
formation, 283 compounds, 419
lattice structure, 273 degradation, 422
light absorption, 283 drugs, 418
molten aluminum, 275 hydroxyl radicals, 421
structural defects, 280 IBU, 421
surface, 277 industrial applications, 419
Ti-H, 284 non-complete mineralization, 422
Oxygen-containing functional groups optimum conditions, 419
(ACOOH), 224, 336, 444 PhACs, 419
Oxyhydroxides, 256 photocatalyst, 419
PIWW, 422
post-filtration step, 422
P real-life wastewater samples, 422
PANI and BF/PANI composite, 174 solar irradiation, 419
PANI/BaTiO3 composites, 196 solvothermal method, 419
PANI/carbon fibre, 187 TiO2, 421, 422
PANI-coated films, 173 toxicity assessment, 422
PANI-coated polyester fabric, 174 toxicity reduction, 423
PANI-coated transparent thin films, 173 Pharmaceutically active compounds
PANI/fabric composites, 174 (PhACs), 419
PANI/graphene hybrid films, 193 Phase engineering, 600, 601
PANI/graphene/amine functionalized Phase polymerization technique, 177
MWCNT hybrid thin film, 189 Phenanthrene, 310, 344
PANI/MWCNT nanocomposites, Phenine nanotubes (pNT), 324
187, 189–191 Photocatalysis, 140, 365, 371, 378, 591, 594,
PANI-TiO2 hybrid-coated cotton fabric, 174 598, 603, 605, 608
PANO nanofibre-based functional cotton, 174 ammonium hydroxide (NH4OH), 10
Para-toluene sulfonic acid (PTSA)-doped anodization process, 7, 8
polyaniline, 172 applications
Index 625

antimicrobial, 22, 23 Photocatalysts


CO2 reduction, 17 heterojunctions (see Heterojunction
dye degradation, 11, 12 photocatalysts)
heavy metal reduction, 19–21 Photocatalytic activity, 114, 300, 301, 309
nitrogen (N2) fixation, 18–20 Photocatalytic antibacterial disinfection, 607
plastic degradation, 14–16 Photocatalytic applications, 114
p pollutant degradation, 12–14 CO2 reduction, 77, 78
challenges, 367 H2 production, 79–81
electrospinning, 8, 9 metal oxide-based semiconductor, 74
energy consumption, 1 water remediation, 74–76
energy crisis, 23 Photocatalytic composite materials, 541
environment, 1 Photocatalytic decomposition, 301, 302,
environmental issues, 23 311, 538
hierarchical nanostructures (see Photocatalytic degradation, 152, 312, 354
Hierarchical nanostructures) MB, 225
hydro-/solvo-thermal synthesis, 5, 6 MO, 225, 226
limitations, 367 organic pollutants, 223
mechanics, 2–4 photogenerated hole, 226
mechanism, 2 RhB, 224, 226, 227
microemulsion method, 10 Photocatalytic degradation efficiency, 127
microemulsion technique, 10 Photocatalytic degradation mechanism, 303
nanostructures, 24 Photocatalytic degradation process, 369
organic pollutants, 365 Photocatalytic degradation reaction, 86
oxidation, 2 Photocatalytic destruction, air pollutants
pechini method, 9 mesoporous materials, 534–539
pollutants, 1, 2 microporous zeolites, 528–533
precipitation process, 6 semiconductor, 540
renewable solar energy, 24 TiO2 photocatalysts, 524
semiconductors, 2 TiO2/porous catalytic adsorbents, 523
sol-gel synthesis, 5 TiO2-supported photocatalysts,
sonochemical method, 10 525–526
spectral irradiance, 367 Photocatalytic disinfection, 445, 446
TiO2, 2 Photocatalytic dye degradation, 592,
ultrasonication, 10 603–605, 608
UV/visible light, 2 Photocatalytic dye degradation advancement
water droplets, 10 adsorption capacities, 415
Photocatalysis mechanism, 546, 547 anthraquinone structure, 412
Photocatalysis process, 366, 373 biocompatibility property, 412
adsorption, 403 CB, 413
organic pollutants removal, 402 dye molecule structure, 412
photocharge carriers, 403 dye wastewater utilization, 412
pollutant oxidation, 405 electrostatic interactions, 415
prematured technology, 403 methylene blue solution, 413
research, 403 natural sunlight, 412
solar energy, 426 non-metal species, 413
technical challenges, 402 pollution degradation efficiency, 415
TiO2, 402 powder-type recovery, 413
value-added recovery materials, 402 solution medium, 414
water-splitting process, 405 TiO2 photocatalysts, 412
Photocatalyst, 371, 374 UV irradiation, 413
literature, 370 Photocatalytic efficiency, 129, 349, 539, 548
magnetic material, 370 Photocatalytic exercises, 493
α–Fe2O3 nanorods, 371 Photocatalytic experiment, 492
Photocatalyst systems, 107, 309 Photocatalytic H2 generation mechanism, 597
626 Index

Photocatalytic hydrogen generation Photogenerated carrier separation, 235


technology, 485 Photogenerated charge carriers, 239
Photocatalytic hydrogen production Photogenerated charges, 285
carbon-based materials, 228 Photogenerated electrons, 114, 218, 223
CdS, 242 Photogenerated holes (h+), 86
charge carrier separation, 229 Photogenerated oxidation-reduction
charge transfer and separation, 228 reactions, 114
photogenerated charge carriers, 227 Photo-induced electrons, 463
rapid charge recombination, 228 Photoinduced redox centre, 549
rGO-TiO2 composite, 229 Photoluminescence, 143
semiconductor photocatalyst, 227 Photon energy, 19
TiO2 nanotubes, 228 Photooxidation capability, 349
W2C@C core/shell, 228 Physical interactions, 331
Photocatalytic material, 87, 113 Physico-chemical properties, 322, 325, 337
Photocatalytic mechanism, 311 Pickering emulsion approach, 103
Photocatalytic memory effect, 2 Plasma-based approach, 285
Photocatalytic nanomaterials, 301 Plasma-enhanced chemical vapor deposition
Photocatalytic NOx removal, 444, 445 (PECVD), 436, 572, 573
Photocatalytic oxidation (PCO), 522 Plasma-enhanced heterostructure
Photocatalytic oxidation system, 523 photocatalyst, 143
Photocatalytic performance, 371 Plasma treatment, 274
Photocatalytic performance analysis Plasmonic composites
COD, 409, 410 Ag/Ag2MoO4 composite, 91
gas chromatography, 408 Au-CuS-TiO2 ternary composite, 92
HPLC, 405–408 co-polymerization method, 91
mass spectrometry, 408 enhanced photocatalytic activity, 91
parameters, 405 oxytetracycline degradation, 92
spectrophotometer, 408, 409 semiconductor material, 90
TOC, 410 SPR, 90, 92
Photocatalytic procedures, 137 visible light irradiation, 90
Photocatalytic process, 113, 348, 522 ZnO surface, 91
Photocatalytic properties, 118 Plasmonic metal, 119
Photocatalytic reaction, 308, 370 Plasmonic photocatalysis, 115
Photocatalytic reactor, 425 Plasmonic photocatalytic materials
Photocatalytic reduction, 154, 155, 226, 447 Ag–metal oxides, 121–123
Photocatalytic systems, 312 Au–metal oxide, 123–124
Photocatalytic water splitting, 227, 449, g-C3N4-based, 128
502, 503 graphene and noble metals, 126–128
Photocatalyts synthesis LSPR probing, 121
Aurivillius compounds, 142 metal nanoparticles, 120
chemical precipitation, 144 photocatalytic reaction, 121
hydrothermal/solvothermal method, Pt–metal oxide, 124–125
142, 143 Plasmonic photocatalytic systems, 120
procedure, 141 Plasmon-induced bacterial disinfection, 156
sol-gel synthesis, 143 Plastic degradation, 14–16
Photochemical smog, 521 PM pollutants, 522
Photo-conversion efficiency, 118 p-n heterojunction composites
Photo-decomposition, 493, 532 Ag/Ag2O/PbBiO2Br, 94
Photodegradation, 355, 440 BiOI/La2Ti2O7 composite, 93
Photodegradation mechanism, 303 charge carrier recombination, 94
Photodeposition method, 603 charge recombination thereby, 93
Photodissolution effect, 376 ciprofloxacin degradation, 94
Photoelectrochemical reaction, 115 CoO@MnCo2O4 hybrid, 93
Photo-excited electrons, 548 CuS/BiVO4 heterostructure, 94
Index 627

hierarchical architecture, 93 PPY-coated carbon nanotubes, 191


isoenergetic electron transfer, 94 PPY-coated fabrics
NGQDs, 94 electrical conductivity, 174
NIR light irradiation, 94 EM wave absorption/reflection, 176
photoexcited charge carrier’s EMI SE, 177
recombination, 93 sodium p-toluenesulfonate-doped, 174
semiconductors, 92 surface microdissolution, 176
PnHMAg samples, 182, 184 PPY-MWCNT-Ag composites, 191
p-nitrophenol degradation, 152 Precipitation method, 68, 69
PNP-sensitized semiconductors, 121 Premature pollution-related deaths, 524
Pollutants, 402 Preparedepoxy-based hybrid composites, 202
conventional treatment systems, 411 Primary air pollutants, 521
external matters/energies, 455 Pristine graphene, 326, 348
heavy metal, 424 Pristine photocatalysts, 312
impacts, drinking water, 411 Pristine unary photocatalyst, 312
organic dye, 411, 412 Process industry wastewater
pesticides, 415–418 (PIWW), 421
pharmaceutical, 420 Proton implantation method, 278
photocatalytic dye degradation, 414 Prototype photocatalytic reactor, 378
Pollution-causing systems, 33 Pseudo-first-order kinetics, 339, 349
Pollution-free environment, 33 Pseudo-second-order kinetics, 342, 357
Poly-and per-fluoroalkyl substances Pt-Cu decorated pyridine-like nitrogen-doped
(PFAS), 368 graphene, 339
Poly(3,4-ethylenedioxythiophene) Pt–metal oxide
(PEDOT), 172 aerobic oxidation, 124
Poly-amidoamine dendrimer, 347 CH4 formation, 125
Polyaniline, 193 intraband/interband transition, 125
Polyaniline hollow microspheres methane production, 124
(PnHM), 182 photocatalytic activities, 125
Polyaniline microtubes, 181 photocatalytic performance, 125
Polyaniline-TiO2-γ-Fe2O3 nanocomposite, 197 photogenerated electron-hole, 126
Polybrominated diphenyl ethers (PBDEs), 343 semiconductor surface, 125
Polychlorinated biphenyls (PCBs), 343 Pt-supported anatase, 125
Polychlorinated dibenzo-p-dioxins, 343 Pulsed laser ablation, 277
Polycyclic aromatic hydrocarbons (PAHs), Pure metal nanoparticles (NPs), 121
343, 368
Polyethyleneimine cross-linked GO, 348
Polymer, 491 Q
Polymer-based nanomaterials, 569, 570 QE decrement, 548
Polymer-based nanoparticles, 578 QE enhancement, 546
Polymer-like semiconductor materials, 491 QSPR modeling approach, 344
Polymer-type semiconductors, 491 Quantitative structure–property relationship
Polyoxotungstate/TiO2 photocatalysts, 306 (QSPR), 344
Polypropylene (PP), 15 Quantum dots (QD), 584
Polypyrrole-chitosan composite films, 178 Quantum efficiencies, 425, 557, 559
Polypyrrole-coated nonwoven textiles, 178 Quartz microbalance (QCM), 582
Polyvinyl borate (PVB), 16 Quaternary photocatalytic systems, 311
Polyvinylpyrrolidone (PVP), 73, 504
Positive dyes (MB), 342
Post-filtration step, 425 R
Potassium hydroxide (KOH), 6 Reactive oxygen species (ROS), 249, 445,
Powder photocatalysts, 425 551, 556
PPY/carbon fibre composite, 177 generation, 366
PPY-coated carbon, 193 photocatalysis process, 366
628 Index

Recyclable photocatalysts Semiconducting metal oxides, 36, 592


cost, 369 Semiconducting nanocomposites, 602
magnetic separation, 369, 370 Semiconducting nanomaterials, 300
photocatalysts, 369 Semiconducting natured nanoparticles, 457
recycling process, 369 Semiconducting photocatalyst materials, 465
Recycling process, 369, 370, 379 Semiconductor, 2, 233, 591, 598, 599
Red phosphorus (RP), 458 Semiconductor-carbon composites
Redlich–Peterson (R–P) isotherm model, 338 carbon sphere-based, 106
Reduced graphene oxide (RGO), 96, 220, 223, CQDs, 102, 104–105
225, 326, 600, 604, 605 semiconductor-graphene, 104
Reduced graphene oxide/MO (rGO/MO) Semiconductor-CQDs composites
heterojunction, 48–50 Bi2MoO6 nanosheet surface, 104
Reduced interplanar distance, 279 BiOI nanosheets, 104
Reduction-oxidation reactions, 2 degradation pathway, 105
Reflectance spectroscopy (DRS), 554 photocatalytic applications, 104
Regulamentation, 260 photocatalytic degradation, 105, 106
Rehybridization, 323 photocenter, 104
Relative permittivity, 171 plasma effect, 105
Remedial technology, 321 Semiconductor-graphene composites
Removal of metal ions, 33 CeM/GO nanocomposite, 103
Resistive-type chemical sensor, 582 charge recombination, 104
Response surface methodology (RSM), 13 GO/LaVO4 nanocomposite, 103
Retention index, 408 hybrid photocatalytic material, 103
Reverse flow reactor (RFR), 353 N-GQDs, 103
Reverse osmosis membranes, 350 photocatalytic performance, 103
RGTO photocatalytic activity, 307 RGO-CdS/ZnS composites, 103
Rhodamine B (RhB), 127, 227, 284, 412 Semiconductor material, 114
ROS scavengers, 425 Semiconductor-metal composites
Rushton turbines, 252 photocatalytic performance, 89
Ruthenium-complex hybrid plasmonic, 91
photocatalysts, 501 Schottky barrier, 89, 90
Semiconductor metal oxides, 36
Semiconductor MoS2, 594
S Semiconductor photocatalysis, 217, 233, 546
Scanning electron microscope (SEM) development, 87
images, 371 photocatalytic process, 87
Schottky barrier composites, 89, 90 photoinduced electron-hole charge
Schottky heterojunction, 442 separation, 87
Schottky junction, 35, 90, 123, 434 Semiconductor photocatalyst, 218, 227, 307
Schottky junction barrier Semiconductor precursor, 438
electron-hole pair recombination, 117 Semiconductor-semiconductor composites
metal-semiconductor interfaces, 117 heterojunction construction, 92
n-type semiconductor, 117 p-n heterojunction, 92–95
schematic, 117 Z-scheme heterojunction (see Z-scheme
Seawater desalination, 351 heterojunction composites)
Secondary air pollutants, 521 Semiconductor wrapping photocatalyst
Secondary shielding mechanism, 168 GP synthesis, 219, 220
Selected area electron diffraction (SAED), 279 H2 production, 218
Selected GR-based photocatalysts, 440 rGO-wrapped TNFs, 220
Selectivity, 523, 526, 528 TiO2 core/shell microspheres
Self-assembly prepared method, 178 encapsulation, 218, 219
Semi-coke-supported TiO2-rGO wrinkled graphene-wrapped TiO2
nanocomposite, 444 nanotubes, 222, 223
Semiconducting materials, 597, 601, 607 Semi-metallic graphene, 229
Index 629

Sensing Solar energy harvesting, 113


mechanism of sensing, 571 Solar light, 66
performance, 567 Solar radiation, 113
Sensor performance, 579 Solar spectrum, 113, 114, 425
Sensors Sol-gel method, 438
biological sensors, 571 Sol-gel process, 576
carbon-based, 571 Sol-gel synthesis, 5, 143
components, 565 Solid-state synthesis method, 578
environmental applications, 565 Solution precursor plasma spray
environmental health monitoring, 569 (SPPS), 555
gas-sensing device, 570 Solvent‐exfoliated graphene/P25 TiO2
micro-sized sensing particles, 565–566 nanocomposite photocatalyst, 443
optical sensors, 566 Solvo-/hydrothermal methods, 592–595, 608
sensing mechanism, 570 Solvothermal method, 70, 71, 275
simple sensors, 570 Spark plasma sintering technique, 580
voltammetric, 571 Spectrophotometer, 408, 409
Shielding efficiency (SE), 168, 170 Spherical/nanodot-loaded semiconductors, 603
Shielding material, 205 Spinning disk reactor (SDR), 250, 252
Shielding mechanism, 190 SPR-induced localized electric field, 126
Si Quantum Dot (QD)-assisted chemical S-scheme mechanism, 509
etching, 278 Stability, 226, 229
SiC thermal deposition, 437 Strain HK-1, 476
Silicon-based nanomaterials, 569 Sulfamethazine degradation, 407
Silver-incorporated ICPs Sulfonamides, 334
AgNWs, 181 Sulfur nitrides, 488, 489
frequency range, 184 Sulphur oxides (SOx), 521
HPPy, 184 Sun energy, 486
PAni composite, 181 Sunlight-derived degradation, 126
PDMS-40, 180 Superbugs, 85
PnHM, 182 Superior photocatalytic efficiency, 106
polyaniline microtubes, 181 Surface area, 8, 13, 15, 23, 24, 390
PPy/Ag/PET nonwoven composite, 184 Surface disorders, 283, 284
precursor concentration, 181 Surface enhanced Raman spectroscopy
shielding effectiveness, 178 (SERS), 126
Simple sensors, 570 Surface functionalization
Single-bilayer ITO/V-doped TiO2 sample, 507 fluoride adsorption capacity, 395, 396
Single-metal nitride fluoride, 496, 497 Ti-O-based materials, 396
cobalt system, 498 TNT and FZTNT, 395
magnesium system, 498 Surface grafting
Single-walled carbon nanotubes (SWCNTs), Cu2+ grafting, 554
324, 573, 574 Cu2+-grafted ZnO, 554
SiO4 tetrahedral, 528 DRS, 554
SLSDPPy-hybrid carbon composites, 193 IFCT, 554
Smaller bandgap, 65 metal oxide nanocluster interface, 553
SnO2/ZnO hierarchical nanostructures, 577 oxidation reaction, 553
SnS2 nanosheets, 460–462 pristine ZnO, 554
SnS2 surface, 461 QE, 554
SnS2/n-propylamine precursor, 460 VB, 554
Soil antibiotic-resistant bacteria, 355 visible light activity, 554
Soil contaminants, 355 Surface modification, 146, 337, 548, 550
Soil microbes, 356 Surface plasma resonance, 548
Soil reclamation, 321, 357 Surface plasmon (SP), 116
Soil remediation, 257, 258, 261 Surface plasmon resonance (SPR), 34, 41, 90,
Soil-nZVI mixing process, 256 92, 128, 129
630 Index

Surface plasmon resonance-induced radiative TiO2 impetus, 486


transfer processes, 116 TiO2 nanorod (TNR), 506
Surface sensitization, 233 TiO2 nanostructures, 37, 40–43, 46, 53
Surface-enhanced Raman scattering TiO2 photocatalyst, 425, 522
(SERS), 120 TiO2 semiconductor, photocatalysts, 525, 526
Surface-interface properties, 33, 34, 36, 38, 45 TiO2/graphene oxide nanocomposites, 447
Sustainability integration, 321 TiO2/rGO/WO3 nanocomposite, 445
Sustainable processes, 402 TiO2-Ag heterojunction, 41, 42
Sustainable solar energy storage/ TiO2-Ag2O heterojunction, 38, 40
conversion, 233 TiO2-Au heterojunction, 42, 43
Synthesis protocol, 559 TiO2-based photocatalysis, 424
Synthesised materials, 449 applications, 267
Synthesized H-doped black titania, 119 basic mechanism, 267
Synthesized rGO-wrapped Cu2O spheres, 226 TiO2-based photocatalysts, 522, 525
Synthesized three-dimensional TiO2-Cu heterojunction, 44, 45
photoactive GP, 224 TiO2-CuO heterojunction, 38, 39
Synthetic carbon allotropes, 433 TiO2-incorporated zeolites (TiO2/zeolite), 530
Synthetic dyes, 411 TiO2-MnO2 heterojunction, 37
Synthetic nanocarbon allotropes, 323 TiO2-Pt heterojunctions, 43, 44
TiO2-RGO (RGTO) aerogel photocatalyst, 307
TiO2-rGO photocatalysts, 445
T TiO2-supported catalysts, 522
TBOT controlled hydrolysis, 218 TiO2-ZnO heterojunction, 39–41
Teflon-lined stainless steel, 69 Ti-OH bonding, 281
Temperature monitoring, 567 Ti-oxide-supported zeolite catalyst, 539
Ternary nanocomposite-based Titanate nanostructures, 386, 388, 390, 391, 397
photocatalyst, 157 biosorbents, 384
Tetrabutyltitanate (TBT), 71 electrodialysis, 384
Tetracycline (TC), 88, 336 fluoride water filters, 385
Tetracycline degradation, 153 inorganic pollutants, 383
Tetracyclines, 332 nanoadsorbents, 385
Tetrasulfurtetranitride, 489 nanostructures, 385
Textile-coated conducting polymers, 172 solid material, 384
Thermal stability, 229, 498 Titanate nanotube (TNT), 395
Thermal treatment Titanium dioxide (TiO2), 522
ambient argon treatment, 272, 273 Titanium nitride (TiN), 505
ambient hydrogen–argon treatment, Titanium nitride-vanadium nitride (TiN-VN)
271, 272 nanohybrids, 500
ambient hydrogen–nitrogen treatment, 272 Titanium oxide-based mesoporous
argon–nitrogen treatment, 273 catalysts, 538
high-pressure hydrogen treatment, 269 Titanium oxynitride (TiOxNy), 506
low-pressure hydrogen treatment, 270 TMNs characteristics, 497
TiO2 transforms, 269 TNCuPc/TiO2, 75, 76
Thermodynamic considerations, 255 Total conductivity, 171
Three-dimensional graphene foam Total electromagnetic interference shielding
(3DGF), 582 effectiveness (SET), 170
Ti(IV) nanocluster, 557 Total organic carbon (TOC), 410
Ti/zeolite-based photocatalysts, 530 Toxic compounds, 524
Ti-MCM-41 material, 534 Toxic gas sensor
TiN/CNT hybrid synthesis, 495 carbon-based nanostructure, 580, 581
TiN@GNS hybrids, 495 humidity sensor, 582, 583
TiO2 catalysts, 527 metal oxide-based nanomaterials, 579, 580
TiO2 core/shell microspheres, 218, 219 research activity, 579
TiO2 crystal structure, 120 sensor properties, 579
Index 631

Toxic pollutants, 411 UV irradiation, 224


Traditional wastewater treatment UV light absorption, 406
techniques, 423 UV light illumination, 557
Transition-metal cations, 73 UV light irradiation, 306
Transition metal nitrides (TMNs), 494, 495 UV-A irradiation, 416
Transition metal-semiconductor, 89 UV-vis diffuse reflection spectra, 224, 409
Transmission electron microscope (TEM), UV-visible light, 309, 310
376, 558
Triazines, 339
Triethylenetetramine, 348 V
Trisodium citrate, 145 Valance band (VB), 115, 128, 283, 300,
2,4-Dichlorophenoxyacetic acid (DPA), 307 311, 555
2-Amino-3-chloro-1,4-naphthoquinone-GO Valence band maximum (VBM), 118
(NQ-GO), 474, 476, 477 Vanadium nitride (VN), 495, 500
2-Aminoanthraquinone-GO (AQ-GO), 474 Vapor collection method, 352
2-Chlorodibenzo-p-dioxin (2-CCD), 309 Variable bandgap energies, 144
2D carbon allotropes, 323 Vibrating sample magnetometry (VSM)
2D covalent organic frameworks (COFs), 472 analysis, 373
2D metal-organic frameworks (2D MOFs), 73 Visible light, 301, 311, 458, 604, 605, 607
2D nanosheet-metal (2D-0D) heterojunction Visible light absorption, 237, 240, 524
photocatalysts, 50–53 Visible light-active catalysts, 558
2D nanosheet-metal oxide (2D-0D) Visible light-active photocatalysts, 546
heterojunction photocatalysts Visible light bandgap semiconductors, 138
g-C3N4/MO, 46, 47 Visible light catalytic reduction, 492
GO/MO, 47–49 Visible light-driven photocatalyst, 591
properties, 45 Visible light-driven photocatalysts, 547, 548
rGO/MO heterojunction, 48–50 Visible light-driven photocatalytic
2D nanosheets, 34 water splitting, 601
2D structure, 593, 597 Visible light excitation, 559
2D structure materials, 460 Visible light irradiation, 129, 599, 606, 607
2D/2D nanocomposites, 154 Visible light photocatalytic ability, 558
2D-0D heterojunction photocatalysts Visible light photocatalytic activity, 556
2D nanosheet-metal (2D-0D), 50–53 Visible light plasmonic photocatalysts, 126
2D nanosheet-metal oxide (see 2D VOC degradation, 553
nanosheet-metal oxide (2D-0D) VOCs photocatalytic destruction, 533
heterojunction photocatalysts) VOCs photocatalytic oxidation, 533, 538
catalytic applications, 45 Volatile organic compounds (VOC), 322, 353,
physicochemical properties, 45 354, 545
structures, 45, 46 air pollutants, 521
2D-2D heterojunction, 101 NOx, 522
2-Propanol degradation, 555 photochemical oxidation, 522
Voltammetric sensors, 571

U
Ultrasonic and hydrothermal (HU) W
method, 458 Wastewater treatment, 39, 40, 47, 53, 137, 329
Ultrasonication, 276 Wastewater treatment plants (WWTPs), 369
Ultrathin graphene materials, 449 Water applications, 248
Ultrathin S-scheme heterojunction, 508 Water cleaning technologies, 401
Ultraviolet (UV), 217 Water pollutant treatment
Uncontrolled chemicals, 299 photocatalysis, 402
US Environmental Protection Agency Water pollutants, 329
(EPA), 415 Water pollution, 148, 154, 605
UV illumination region, 122 Water remediation, 74–76, 81
632 Index

Water splitting, 79, 80, 285, 290, 405, 511, Zinc (II) 2,5-dihydroxy-1,4-­
591, 593, 597, 598, 600, 601, 603 benzenedicarboxylate MOF, 73
Water treatment industry, 401 ZnIn2S4-In2O3, 77
Wedge-shaped/stilt-shaped pores, 332 ZnO hierarchical nano-/microspheres
WO3 hierarchical nanostructure, 71 (ZHN), 303
WO3/g-C3N4 2D/2D heterojunctions, 508 ZnO-MoS2-RGO nanocomposite, 604
WO3/ZnO nanocomposite, 309 ZnO nanostructures, 39, 41, 46, 48, 49,
World Health Organization (WHO), 299 53, 54
Wrinkled graphene-wrapped TiO2 nanotubes, ZnO/TiO2/Au ternary nanostructure, 310
222, 223, 228 ZNR-WO3 photocatalyst, 309
Zr-based metal organic framework
(MOF), 338
X Z-scheme-based MoS2
X-band frequency, 186 nanocomposites, 601–602
X-ray absorption near-edge structure Z-scheme-based photocatalytic activity, 150
(XANES), 549 Z-scheme heterojunction, 35, 150, 152
X-ray photoelectron spectroscopy (XPS), 280, Z-scheme heterojunction composites
281, 393 advantages, 95
XRD process, 496 all-solid-state, 95–98
band structure semiconductors, 100
charge recombination rate, 102
Y charge transfer mechanism, 95
Young’s modulus, 437 direct/mediator-free, 98–100
Yttrium-immobilized-GO-alginate electron-hole separation, 101
hydrogel, 336 g-C3N4/TiO2, 102
In2S3/InVO4 heterojunction, 101
photocatalytic applications, 95
Z TiO2/SnNb2O6 heterojunction, 101
Zeolite-based photocatalysts, 528, 530 Z-scheme mechanism, 152
Zeolites, 528 Z-scheme mes-Sn3O4/g-C3N4 heterostructure
Zeo-type materials, 528 composite, 98
Zero-dimensional nanocluster Z-scheme photocatalyst, 311
preparations, 569 Z-scheme photocatalysts transport, 230
Zero-dimensional structures, 569 Z-scheme photocatalytic system, 149

You might also like