Download as pdf or txt
Download as pdf or txt
You are on page 1of 382

9 Series: Sustainable Energy Developments

9
Advanced Oxidation Technologies (AOTs) or Processes (AOPs) are relatively new and innovative

Litter, Candal & Meichtry


technologies to remove harmful and toxic pollutants. The most important processes among
them are those using light, such as UVC/H2O2, photo-Fenton and heterogeneous photocatalysis
with TiO2. These technologies are also relatively low-cost and therefore useful for countries
under development, where the economical resources are scarcer than in developed countries.

This book provides a state-of-the-art overview on environmental applications of Advanced


Oxidation Technologies (AOTs) as sustainable, low-cost and low-energy consuming treatments
for water, air, and soil. It includes information on innovative research and development on TiO2
photocatalytic redox processes, Fenton, Photo-Fenton processes, zerovalent iron technology,
and others, highlighting possible applications of AOTs in both developing and industrialized

Advanced Oxidation Technologies


countries around the world in the framework of “A crosscutting and comprehensive look at
environmental problems”.

The book is aimed at professionals and academics worldwide, working in the areas of water
resources, water supply, environmental protection, and will be a useful information source for
decision and policy makers and other stakeholders working on solutions for environmental
problems.

SUSTAINABLE ENERGY DEVELOPMENTS – VOLUME 9 ISSN 2164-0645

The book series addresses novel techniques and measures related to sustainable energy
developments with an interdisciplinary focus that cuts across all fields of science, engineering
and technology linking renewable energy and other sustainable materials with human society. It
addresses renewable energy sources and sustainable policy options, including energy efficiency
and energy conservation to provide long-term solutions for key-problems of industrialized, Advanced Oxidation Technologies
developing and transition countries by fostering clean and domestically available energy and,
concurrently, decreasing dependence on fossil fuel imports and reducing greenhouse gas Sustainable solutions for environmental treatments
emissions. Possible applications will be addressed not only from a technical point of view, but
also from economic, financial, social, political, legislative and regulatory viewpoints. The book
series aims to become a state-of-the-art source for a large group of readers comprising different
stakeholders and professionals, including government and non-governmental organizations and
institutions, international funding agencies, universities, public health and energy institutions,
and other relevant institutions.

SERIES EDITOR: Jochen Bundschuh

Editors: Marta I. Litter, Roberto J. Candal & J. Martín Meichtry

an informa business
ADVANCED OXIDATION TECHNOLOGIES –
SUSTAINABLE SOLUTIONS FOR ENVIRONMENTAL TREATMENTS
This page intentionally left blank
Sustainable Energy Developments

Series Editor

Jochen Bundschuh
University of Southern Queensland (USQ), Toowoomba, Australia
Royal Institute of Technology (KTH), Stockholm, Sweden

ISSN: 2164-0645

Volume 9
This page intentionally left blank
Advanced Oxidation Technologies –
Sustainable Solutions for
Environmental Treatments

Editors

Marta I. Litter
Remediation Technologies Division, Environmental Chemistry Department,
Chemistry Management, National Atomic Energy Commission, Buenos Aires;
National Scientific and Technique Research Council (CONICET);
Institute of Research and Environmental Engineering,
National University of General San Martín, Argentina

Roberto J. Candal
National University of General San Martín, School of Science and Technology;
National Scientific and Technique Research Council (CONICET),
Institute of Physical chemistry of Materials, Environment and Energy (INQUIMAE),
University of Buenos Aires, Buenos Aires, Argentina

J. Martín Meichtry
Remediation Technologies Division, Environmental Chemistry Department,
Chemistry Management, National Atomic Energy Commission, Buenos Aires;
National Scientific and Technique Research Council (CONICET);
Chemistry Department, Buenos Aires Regional, National Technological University,
Buenos Aires, Argentina
Cover photo
The cover photo (by Malena Bystrowicz, malebystro@yahoo.com.ar, 2009) depicts the aspect of
the Riachuelo river, in the periphery of Buenos Aires City (Argentina), belonging to the Matanza-
Riachuelo River Basin, one of the top 10 worst polluted places in the world. About 15,000
industries dispose of waste into the river and numerous chemical plants are responsible for more
than one third of the pollution. The deposited levels of zinc, lead, copper, nickel and chrome on
the riverbanks of the Riachuelo exceed the recommended levels. Approximately 60% of around
20,000 reside on the peripheries of the river live in zones considered inappropriate for humans.

CRC Press/Balkema is an imprint of the Taylor & Francis Group, an informa business

© 2014 Taylor & Francis Group, London, UK


Typeset by MPS Limited, Chennai, India
Printed and bound in The Netherlands by PrintSupport4U, Meppel

All rights reserved. No part of this publication or the information contained herein may be
reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic,
mechanical, by photocopying, recording or otherwise, without written prior permission from
the publishers.
Although all care is taken to ensure integrity and the quality of this publication and the
information herein, no responsibility is assumed by the publishers nor the author for any
damage to the property or persons as a result of operation or use of this publication and/or
the information contained herein.

Library of Congress Cataloging-in-Publication Data


Advanced oxidation technologies : sustainable solutions for environmental treatments /
editors, Marta I. Litter, Remediation Technologies Division, Environmental Chemistry
Department, Chemistry Management, National Atomic Energy Commission, Buenos Aires,
Roberto J. Candal, National University of General San Martín, School of Science and
Technology, J. Martín Meichtry, Remediation Technologies Division, Environmental
Chemistry Department, Chemistry Management, National Atomic Energy Commission,
Buenos Aires.
pages cm. — (Sustainable energy developments, ISSN 2164-0645 ; volume 9)
Includes bibliographical references and index.
ISBN 978-1-138-00127-5 (hardback)
1. Sewage—Purification—Oxidation. 2. Oxidation—Environmental aspects.
3. Environmental chemistry—Technique. 4. Sustainable engineering. I. Litter, Marta I.,
editor of compilation. II. Candal, Roberto J., editor of compilation. III. Meichtry, J. Martín
(Jorge Martín), editor of compilation.
TD758.A38 2013
628.3’5—dc23
2013043723
Published by: CRC Press/Balkema
P.O. Box 11320, 2301 EH Leiden, The Netherlands
e-mail: Pub.NL@taylorandfrancis.com
www.crcpress.com – www.taylorandfrancis.com
ISBN: 978-1-138-00127-5 (Hardback)
ISBN: 978-1-315-77765-8 (eBook PDF)
About the book series

Renewable energy sources and sustainable policies, including the promotion of energy efficiency
and energy conservation, offer substantial long-term benefits to industrialized, developing and
transitional countries. They provide access to clean and domestically available energy and lead to
a decreased dependence on fossil fuel imports, and a reduction in greenhouse gas emissions.
Replacing fossil fuels with renewable resources affords a solution to the increased scarcity and
price of fossil fuels. Additionally it helps to reduce anthropogenic emission of greenhouse gases
and their impacts on climate change. In the energy sector, fossil fuels can be replaced by renewable
energy sources. In the chemistry sector, petroleum chemistry can be replaced by sustainable or
green chemistry. In agriculture, sustainable methods can be used that enable soils to act as carbon
dioxide sinks. In the construction sector, sustainable building practice and green construction can
be used, replacing for example steel-enforced concrete by textile-reinforced concrete. Research
and development and capital investments in all these sectors will not only contribute to climate
protection but will also stimulate economic growth and create millions of new jobs.
This book series will serve as a multi-disciplinary resource. It links the use of renewable
energy and renewable raw materials, such as sustainably grown plants, with the needs of human
society. The series addresses the rapidly growing worldwide interest in sustainable solutions. These
solutions foster development and economic growth while providing a secure supply of energy.
They make society less dependent on petroleum by substituting alternative compounds for fossil-
fuel-based goods. All these contribute to minimize our impacts on climate change. The series
covers all fields of renewable energy sources and materials. It addresses possible applications
not only from a technical point of view, but also from economic, financial, social and political
viewpoints. Legislative and regulatory aspects, key issues for implementing sustainable measures,
are of particular interest.
This book series aims to become a state-of-the-art resource for a broad group of readers includ-
ing a diversity of stakeholders and professionals. Readers will include members of governmental
and non-governmental organizations, international funding agencies, universities, public energy
institutions, the renewable industry sector, the green chemistry sector, organic farmers and farm-
ing industry, public health and other relevant institutions, and the broader public. It is designed
to increase awareness and understanding of renewable energy sources and the use of sustainable
materials. It also aims to accelerate their development and deployment worldwide, bringing their
use into the mainstream over the next few decades while systematically replacing fossil and
nuclear fuels.
The objective of this book series is to focus on practical solutions in the implementation of
sustainable energy and climate protection projects. Not moving forward with these efforts could
have serious social and economic impacts. This book series will help to consolidate international
findings on sustainable solutions. It includes books authored and edited by world-renowned
scientists and engineers and by leading authorities in economics and politics. It will provide a
valuable reference work to help surmount our existing global challenges.
Jochen Bundschuh
(Series Editor)

vii
This page intentionally left blank
Editorial board

Morgan Bazilian Deputy Director, Institute for Strategic Energy Analysis (JISEA),
National Renewable Energy Lab (NREL), Golden, CO, USA,
morgan.bazilian@nrel.gov
Robert K. Dixon Leader, Climate and Chemicals, The Global Environment Facility,
The World Bank Group, Washington, DC, USA, rdixon1@thegef.org
Maria da Graça Carvalho Member of the European Parliament, Brussels & Professor
at Instituto Superior Técnico, Technical University of Lisbon,
Portugal, maria.carvalho@ist.utl.pt,
mariadagraca.carvalho@europarl.europa.eu
Rainer Hinrichs-Rahlwes President of the European Renewable Energy Council (EREC),
President of the European Renewable Energies Federation (EREF),
Brussels, Belgium; Board Member of the German Renewable Energy
Federation (BEE), Berlin, Germany, rainer.hinrichs@bee-ev.de
Eric Martinot Senior Research Director, Institute for Sustainable Energy Policies
(ISEP), Nakano, Tokyo & Tsinghua University, Tsinghua-BP Clean
Energy Research and Education Center, Beijing, China,
martinot@isep.or.jp, martinot@tsinghua.edu.cn
Veena Joshi Senior Advisor-Energy, Section Climate Change and Development,
Embassy of Switzerland, New Delhi, India, veena.joshi@sdc.net
Christine Milne Leader of the Australian Greens Party, Senator for Tasmania,
Parliament House, Canberra, ACT & Hobart, TAS, Australia

ADVISORY EDITORIAL BOARD

ALGERIA
Hacene Mahmoudi (renewable energy for desalination and water treatment), Faculty of
Sciences, Hassiba Ben Bouali University, Chlef
ARGENTINA
Marta Irene Litter (advanced oxidation technologies, heterogeneous photocatalysis),
Gerencia Química, Comisión Nacional de Energía Atómica, San Martín, Prov. de Buenos
Aires, Argentina & Consejo Nacional de Investigaciones Científicas y Técnicas, Buenos
Aires, Argentina & Instituto de Investigación e Ingeniería Ambiental, Universidad de General
San Martín, Prov. de Buenos Aires
AUSTRALIA
Thomas Banhazi (biological agriculture; sustainable farming, agriculture sustainable energy
solutions), National Centre of Engineering in Agriculture, University of Southern
Queensland, Toowoomba, QLD
ix
x Editorial board

Ramesh C. Bansal (wind, PV, hybrid systems), School of Information Technology &
Electrical Engineering, The University of Queensland, St. Lucia, Brisbane, QLD

Andrew Blakers (solar energy, solar cell photovoltaic technology), Director, Centre for
Sustainable Energy Systems and Director, ARC Centre for Solar Energy Systems, Australian
National University, Canberra, ACT

John Boland (energy meteorology), School of Mathematics and Statistics and Barbara Hardy
Institute, University of South Australia, Adelaide, SA

Dan Cass (climate politics, community energy, environmental movement), Melbourne, VIC

Guangnan Chen (sustainable energy applications in agriculture), Faculty of Health,


Engineering and Sciences, University of Southern Queensland & National Centre for
Engineering in Agriculture, Toowoomba, QLD

Tom Denniss (ocean energy), Oceanlinx Ltd., Macquarie Park, NSW

Peter Droege (renewable energy autonomy and cities, urban energy transition), Institute of
Architecture and Planning, University of Liechtenstein, Vaduz, Liechtenstein & Faculty of
Engineering, University of Newcastle, Newcastle, NSW

Barry A. Goldstein (geothermal energy: regulation and investment attraction for


exploration/production), Energy Resources – Department for Manufacturing, Innovation,
Trade, Resources and Energy, State Government of South Australia, Adelaide, SA

Hal Gurgenci (Enhanced Geothermal Systems; power generation), Director – Queensland


Geothermal Energy Centre of Excellence, The University of Queensland, Brisbane, QLD

Brigitte House (environment movement, social justice and welfare, life coaching, community
development), Melbourne, VIC

Edson Nakagawa CSIRO, Director – Petroleum and Geothermal Research Portfolio,


Australian Resources Research Centre (ARRC), Kensington, WA

Bibhash Nath (geothermal energy, energy, water & pollution behavior), School of
Geosciences, University of Sydney, Sydney, NSW

Klaus Regenauer-Lieb (thermo-hydro-mechanical-chemical reservoir simulation), Director –


Western Australian Geothermal Centre of Excellence, CSIRO Earth Science and Resource
Engineering and School of Earth and Environment, The University of Western Australia and
Curtin University, Perth, WA

Alberto Troccoli (climate and energy/energy meteorology), Weather & Energy Research Unit
(WERU), CSIRO Marine and Atmospheric Research, Canberra, ACT

Matthew Wright (zero emission strategies), Executive Director, Beyond Zero Emissions,
Melbourne, VIC

Talal Yusaf (alternative fuels for IC engines, micro-organism treatment, microalgae fuel –
production and applications), Faculty of Health, Engineering and Science, University of
Southern Queensland, Toowoomba, QLD

AUSTRIA

Roland Dimai (electromobility: intersection green power generation automotive industry;


needs of human sustainable e-mobility), REFFCON GmbH, Dornbirn
Editorial board xi

BELGIUM
Amelia Hadfield (energy security, energy policies), European Affairs & Institute for
European Studies Energy, Vrije Universiteit Brussel (VUB), Brussels
Klaus Rave (wind energy, financing), Investitionsbank Schleswig-Holstein, Kiel, Germany;
Chairman of the GlobalWind Energy Council & Vice President, European Wind Energy
Association (EWEA), Brussels
BRAZIL
Gilberto De Martino Jannuzzi (energy for sustainable development), Center for Energy
Studies (NIPE), University of Campinas (UNICAMP), Campinas, SP
José Goldemberg (biofuels), Universidade de São Paulo, Sao Paulo, SP
Roberto Schaeffer (energy efficiency, renewable energy and global climate change), Federal
University of Rio de Janeiro, Rio de Janeiro, RJ
Geraldo Lúcio Tiago Filho (sustainable hydropower, renewable energy in general), National
Reference Center for Small Hydropower, University of Itajubá, Itajubá, MG
CANADA
Ghazi A. Karim (hydrogen technologies), Department of Mechanical and Manufacturing
Engineering, University of Calgary, Calgary, AB
Xianguo Li (fuel cells, energy and exergy analysis, energy efficiency), Department of
Mechanical Engineering, University of Waterloo, Waterloo, ON
Marc A. Rosen (modeling of energy systems, exergy, district energy, thermal energy storage),
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology,
Oshawa, ON
Erik J. Spek (electric cars, batteries/energy storage), TÜV SÜD Canada, Newmarket, ON
Sheldon S. Williamson (electric and hybrid electric vehicles, automotive power electronics
and motor drives), Department of Electrical and Computer Engineering, Concordia
University, Montreal, Quebec, QC
Laurence T. Yang (green(ing) computing), Department of Computer Science, St. Francis
Xavier University, Antigonish, NS
CYPRUS
Soteris Kalogirou (solar energy and desalination), Department of Mechanical Engineering
and Materials Sciences and Engineering, Cyprus University of Technology, Limasol
DENMARK
Søren Linderoth (fuel cells and hydrogen technologies), Head of Department, Department of
Energy Conversion and Storage, Technical University of Denmark, Roskilde
Kim Nielsen (ocean energy), Ramboll, Virum
EGYPT
Galal Osman (wind energy), Egyptian Wind Energy Association, Cairo
FIJI ISLANDS
Thomas Lynge Jensen (sustainable energy for small islands), UNDP Pacific Centre (PC),
Suva
xii Editorial board

FINLAND
Pertti Kauranen (nanotechnologies for sustainable energy applications), VTT Advanced
Materials, Tampere
FRANCE
Bruno Francois (renewable energy based electrical generators, smart grids), Laboratoire
d’Electrotechnique et d’Electronique de Puissance, Ecole Centrale de Lille, Paris
Sébastien Martinet (batteries for electric and hybrid vehicles), Département Electricité et
Hydrogène pour les Transports, CEA – LITEN/DEHT, Grenoble
Jérôme Perrin (electric vehicles), VP Director Advanced Projects for CO2 , Energy and
Environment, Renault, Guyancourt
GERMANY
Holger Dau (bio-inspired solar fuel production/water splitting/solar H2 ), Department of
Physics, Freie Universität Berlin, Berlin
Claus Doll (hybrid electric vehicles; electric vehicles and mobility concepts; adapting
transport to climate change), Fraunhofer-Institute for Systems and Innovation Research,
Karlsruhe
Hans-Josef Fell (solar and renewable energy), Member of the German Parliament
(1998–2013), Spokesperson on energy for the Alliance 90/The Greens parliamentary group
in the German Parliament (2005–2013), Berlin
Jan Hoinkis (renewable energy for water treatment), Institute of Applied Research, Karlsruhe
University of Applied Sciences, Karlsruhe
Ernst Huenges (geothermal reservoir technologies), Helmholtz-Zentrum Potsdam, Deutsches
GeoForschungsZentrum, Potsdam
Rainer Janssen (bioenergy, biofuels, RE strategies and policies, capacity building and
communication strategies), WIP Renewable Energies, München
Claudia Kemfert (energy economics, RE strategies), Department of Energy, Transportation
and Environment, German Institute for Economic Research (DIW) & Hertie School of
Governance, Berlin
Thomas Kempka (geological CO2 storage), Helmholtz Centre Potsdam, German Research
Centre for Geosciences, Potsdam
Harry Lehmann (sustainability strategies and instruments, climate protection), General
Director, Division I Environmental Planning and Sustainability Strategies, Federal
Environment Agency of Germany, Dessau
Wolfgang Lubitz (bio-inspired solar fuel production/solar H2 ), Max-Planck-Institut for
Bioinorganic Chemistry, Mülheim an der Ruhr
Thomas Ludwig (green(-ing) computing, energy-efficient high-performance computing),
University of Hamburg, Hamburg
Gerd Michelsen (education for sustainability, communication strategies), Institut für
Umweltkommunikation (INFU), Leuphana Universität Lüneburg, Lüneburg
Dietrich Schmidt (pre-industrial developments for sustainable buildings, energy efficiency),
Head of Department Energy Systems, Fraunhofer Institute for Building Physics, Project
Group Kassel, Kassel
Editorial board xiii

Frank Scholwin (biogas/biomethane), Scientific Managing Director, DBFZ Deutsches


Biomasseforschungszentrum GmbH, Leipzig, Germany/University Rostock, Rostock

Martin Wietschel (electromobility), Competence Center Energiepolitik, und Energiesysteme,


Fraunhofer-Institut für System- und Innovationsforschung ISI, Karlsruhe

Wolfgang Winkler (fuel cells), Hamburg University of Applied Sciences,


Forschungsschwerpunkt Brennstoffzellen und rationelle Energieverwendung, Hamburg

GREECE

Eftihia Tzen (water desalination, desalination with renewable energy sources), Wind Energy
Department, Centre for Renewable Energy Sources & Saving, Pikermi

HONG KONG

Dennis Leung (energy conversion and conservation), Department of Mechanical


Engineering, The University of Hong Kong, Hong Kong

Tim S. Zhao ((alcohol) fuel cells, heat/mass transport phenomena), Center for Sustainable
Energy Technology, The Hong Kong University of Science & Technology, Hong Kong

HUNGARY

Jamal Shrair (nanotechnologies for sustainable energies), Department of Electronic Devices,


Budapest University of Technology and Economics, Budapest

INDIA

Rangan Banerjee (energy systems modeling, energy efficiency, renewable energy),


Department of Energy Science and Engineering, Indian Institute of Technology Bombay,
Mumbai

Jens Burgtorf (CDM capacity building: sustainable energy strategies), Director, Indo-German
Energy Programme (IGEN) – Deutsche Gesellschaft für Internationale Zusammenarbeit
(GIZ) GmbH, Bureau of Energy Efficiency, New Delhi

D. Chandrashekharam (geothermal resources in developing countries), Indian Institute of


Technology, IIT Bombay, Mumbai

Shanta Chatterji (electromobility in developing urban cities, public awareness), Chattelec


Vehicles India Ltd & Clean Air Island, Mumbai

Sudipta De (sustainable energy engineering), Mechanical Engineering Department, Jadavpur


University, Kolkata

Arun Kumar (sustainable hydropower), Alternate Hydro Energy Centre, IIT Roorkee,
Roorkee, Uttarakhand

Naveen Kumar (biodiesel) Mechanical Engineering and Head, Biodiesel Research, Delhi
College of Engineering, Delhi

Jayant K. Nayak (passive solar architecture, energy conscious building), Indian Institute of
Technology, IIT Bombay, Mumbai

Ambuj D. Sagar (bioenergy, rural electrification), Vipula and Mahesh Chaturvedi Chair in
Policy Studies, Department of Humanities and Social Sciences, Indian Institute of
Technology, IIT Delhi, New Delhi
xiv Editorial board

INDONESIA
Alessandro Palmieri (sustainable hydropower), The World Bank (Jakarta office), Jakarta
IRELAND
Eoin Sweeney (ocean energy), Ocean Energy Development Unit, Sustainable Energy
Authority of Ireland, Dublin
ISLAND
Guðni Jóhannesson (geothermal powered buildings, low energy systems in buildings),
Director General, Orkustofnun – National Energy Authority, Reykjavík
ITALY
Ruggero Bertani (geothermal power generation), Geothermal Center of Excellence, Enel
Green Power, Rome
Pietro Menga (e-mobility), CIVES, Milan
Gianfranco Pistoia (Li and Li-ion batteries, electric vehicles), Consultant, Rome
JAPAN
Yoichi Hori (electric vehicles, motion control), University of Tokyo, Tokyo
Tetsunari Iida (sustainable energy policies, financing schemes), Executive Director, Institute
for Sustainable Energy Policies (ISEP), Nakano, Tokyo
MEXICO
Sergio M. Alcocer (ocean energy), Instituto de Ingeniería UNAM, Mexico DF
Omar R. Masera (bioenergy), Center for Ecosystems Research, Universidad Nacional
Autónoma de México (UNAM), Morelia, Michoacán
Mario-César Suarez-Arriaga (geothermal reservoirs, numerical modeling of complex
systems), Facultad de Ciencias Físico-Matemáticas, Universidad Michoacana de San Nicolás
de Hidalgo (UMSNH), Morelia, Michoacán
NIGERIA
Adeola Ijeoma Eleri (biogas, sustainable energy solutions), Renewable Energy Department,
Energy Commission of Nigeria, Abuja

NORWAY
Einar Hope (energy economics), Norwegian School of Economics & Business
Administration, Bergen
Ånund Killingtveit (sustainable hydropower), Norwegian University of Science and
Technology (NTNU), Trondheim
Harald N. Røstvik (solar cars, solar buildings), Architect MNAL, Holder of the Sustainability
Professorship Bergen School of Architecture, Sunlab, Stavanger

PARAGUAY
Wolfgang F. Lutz (sustainable energy policies, energy efficiency, renewable energy), Energy
Strategies for Sustainable Development/Estrategias Energéticas para un Desarrollo
Sustentable, Ter Aar, The Netherlands/Asunción
Editorial board xv

POLAND
Antoni Szumanowski (drives for electric and hybrid vehicles), Faculty of Automotive and
Construction Machinery Engineering, Warsaw University of Technology, Warsaw

P.R. CHINA
Ma Jiming (sustainable hydropower), Department of Hydraulic Engineering, Tsinghua
University, Beijing

SAUDI ARABIA
Gary L. Amy (renewable energy for desalination and water treatment), Director, Water
Desalination and Reuse Research Center, King Abdullah University of Science and
Technology (KAUST)
Peter Birkle (geochemistry of geothermal and petroleum reservoirs), Saudi Aramco,
Exploration and Producing Advanced Research Center (EXPO ARC), Geology Technology
Team (GTT), Dhahran
Noreddine Ghaffour (renewable energy for desalination and water treatment), Water
Desalination and Reuse Research Center, King Abdullah University of Science and
Technology (KAUST)
SINGAPORE
Siaw Kiang Chou (energy performance of buildings), Executive Director, Energy Studies
Institute (ESI) & Department of Mechanical Engineering, National University of Singapore
(NUS)
SPAIN
Santiago Arnaltes (wind energy), Wind to Power System, S.L., Getafe (Madrid)
Josep Puig (renewable energy policies and community power), EUROSOLAR Spain,
Barcelona, Catalunya
Guillermo Zaragoza (solar energy and desalination), CIEMAT-Plataforma Solar de Almería,
Almería
SRI LANKA
Ravindra Anil Cabraal (grid-connected and off-grid renewable energy), Director, KMRI
Lanka (pvt) Ltd., a renewable power development company; Consultant, Renewable and Rural
Energy, Colombo; formerly: Lead Energy Specialist at The World Bank,Washington DC
SWEDEN
Prosun Bhattacharya (sustainable energy and water), Department of Land and Water
Resources Engineering, Royal Institute of Technology (KTH), Stockholm
Erik Dahlquist (bio-mass/bio-energy, biomass combustion), Malardalen University, Energy
Engineering, Västerås
Thomas B. Johansson (energy for sustainable development), International Institute for
Industrial Environmental Economics, Lund University, Co-Chair, Global Energy Assessment,
IIASA, Lund
Andrew Martin (membrane distillation for desalination and water purification; environomic
modeling; biomass and municipal solid waste; concentrating solar thermal power),
Department of Energy Technology, Royal Institute of Technology (KTH), Stockholm
xvi Editorial board

Semida Silveira (sustainable energy solutions for development, infrastructure systems,


policies and entrepreneurship for sustainable development), Department of Energy
Technology, Royal Institute of Technology, Stockholm
Maria Wall (energy-efficient buildings), Energy and Building Design, Department of
Architecture and Built Environment, Lund University, Lund
Ramon Wyss (Innovations for sustainable energy systems), Vice President International
Affairs, Royal Institute of Technology (KTH); KTH Energy Platform Coordinator, Stockholm
Jinyue Yan (biofuels, bioenergy), School of Chemical Science and Engineering, Div. of
Energy Processes, Royal Institute of Technology (KTH), Stockholm
SWITZERLAND
François Avellan (hydropower and hydraulic turbomachines), Laboratoire de Machines
Hydrauliques (LMH), École Polytechnique Fédérale de Lausanne (EPFL), Lausanne
Urs Muntwyler (photovoltaics system technology, electric and hybrid vehicles), Photovoltaic
Laboratory, Bern University of Applied Sciences, Engineering and Information Technology,
Burgdorf
Ladislaus Rybach (geothermal energy, heat pumps, EGS), Geowatt AG, Zurich
Robert Stüssi (transport policy, sustainable mobility, electric vehicles), Owner of
Robert.Stussi Mobil (consulting), Portuguese Electric Vehicle Association (President) and
past president of AVERE and WEVA, Lisboa, Portugal and ZUG

TAIWAN
Shih Hung Chan (fuel cells, hydrogen technologies), Fuel Cell Center, Yuan Ze University,
Taipei
S.K. Jason Chang (sustainable public transportation: planning, policy, economy, operation),
National Taiwan University, Department of Civil Engineering, Taipei
Falin Chen (fuel cells, hydrogen technologies), Director, National Taiwan University Energy
Research Centre, Taipei
THAILAND
Wattanapong Rakwichian (renewable energy education, carbon-free cities), Director, Asian
Development Institute for Community Economy and Technology (adiCET), Chiang Mai
Rajabhat University, Chiang Mai
Lisa Schipper (development and adaptation to climate change – policy, science and practice),
Stockholm Environmental Institute, Bangkok
THE NETHERLANDS
Jaco Appelman (green(-ing) computing), Delft University of Technology, Delft
Frances Brazier (green computing), Delft University of Technology, Delft
Rafid al Khoury (geothermal and geological CO2 sequestration modeling), Faculty of Civil
Engineering and Geosciences, Delft University of Technology, Delft
Rob Kool (energy efficiency), NL Agency, Utrecht; Board member of the European Council
for an Energy Efficient Economy (ECEEE) and Chair of IEA DSM & IEA EGRD
Editorial board xvii

TUNISIA
Thameur Chaibi (geothermal water desalination), National Institute for Research in Rural
Engineering Water and Forestry (INRGREF), Tunis
TURKEY
Hamdi Ucarol (electric and hybrid vehicles), Energy Institute, TÜBITAK Marmara Research
Center, Gebze/Kocaeli
UAE
Khaled A. Al-Sallal (low energy architecture), Faculty of Engineering, UAE University,
Al-Ain
UK
AbuBakr S. Bahaj (ocean energy), School of Civil Engineering and the Environment,
University of Southampton, Southampton
Philip A. Davies (renewable and efficient energy systems and their application for water
treatment and agriculture), Sustainable Environment Research Group, School of Engineering
and Applied Science, Aston University, Birmingham
USA
Suresh K. Aggarwal (combustion simulations, renewable fuels), University of Illinois at
Chicago, IL
Ishfaq Ahmad (green computing), University of Texas at Arlington, TX
Said Al-Hallaj (hybrid hydrogen systems, solar water desalination), Chairman/ CEO AllCell
Technologies, LLC, & Department of Chemical Engineering, University of Illinois at
Chicago, Chicago, IL
Kalyan Annamalai (combustion, biomass, animal waste, energy conversion), Texas A&M
University, College Station, TX
Joel R. Anstrom (hybrid and hydrogen vehicles), Director of the Hybrid and Hydrogen
Vehicle Research Laboratory, Larson Transportation Institute, University Park, PA
Ronald Bailey (electric vehicles), Center for Energy, Transportation and the Environment,
University of Tennessee at Chattanooga, Chattanooga, TN
Gary W. Brudvig (bio-inspired solar fuel production/solar H2 ), Department of Chemistry,
Yale University, New Haven, CT
Kirk W. Cameron (green(ing) computing), SCAPE Laboratory and Department of Computer
Science, Virginia Tech, Blacksburg, VA
Daniel Cohn (hydrogen technologies for transportation), Plasma Science and Fusion Center,
Massachusetts Institute of Technology (MIT), Cambridge, MA
Kristin Deason (fuel cells), National Organization Wasserstoff and
Brennstoffzellentechnologie (NOW), Berlin, Germany & SENTECH, Washington DC
Gregory Dolan (methanol fuels), Methanol Institute, Alexandria, VA
James Edmonds (global climate change), Pacific Northwest National Laboratory, Joint
Global Change Research Institute at the University of Maryland, College Park, MD
Ali Emadi (hybrid and plug-in hybrid vehicles), Director, Electric Power and Power
Electronics Center and Grainger Laboratories, Electrical and Computer Engineering
Department, Illinois Institute of Technology (IIT) in Chicago, IL
xviii Editorial board

Andrew Frank (plug-in hybrid electric vehicles), Department of Mechanical and Aerospace
Engineering, University of California, Davis, CA, and CTO of Efficient Drivetrains Inc.
Petra Fromme (bio-inspired solar fuel production/solar H2 ), Department of Chemistry and
Biochemistry, Arizona State University, Phoenix, TX
Vasilis Fthenakis (energy & sustainability, solar energy, renewable energy penetration in the
grid, CAES), PV Environmental Research Center, Brookhaven National Laboratory and
Center of Life Cycle Analysis, Columbia University, New York, NY
Chris Gearhart (fuel cells for transportation), Fuel Cell System Research, Ford Motor
Company, Dearborn, MI
John Golbeck (bio-inspired solar fuel production), Pennsylvania State University, University
Park, PA
Barbara Goodman (sustainable energy technologies for transportation), Center for
Transportation Technologies and Systems, National Renewable Energy Laboratory (NREL),
Golden, CO
James Gover (hybrid electric vehicles), IEEE Fellow Professor of Electrical Engineering,
Kettering University, Flint, MI
Allan R. Hoffman (solar and wind energy), former Deputy Assistant Secretary for Utility
Technologies, former Director Country Studies Program, and former Senior Analyst – all in
the U.S. Department of Energy’s Office of Energy Efficiency and Renewable Energy. Now
visiting Professor for Renewable Energy and Desalination at GORD (Gulf Organization for
Research & Development)/Qatar and Visiting Research Professor at University of
Aberdeen/Scotland. Home: Reston, VA
Iqbal Husain (electric and hybrid vehicles), Department of Electrical & Computer
Engineering, The University of Akron, Akron, OH
Gerald W. Huttrer (geothermal energy), Geothermal Management Company, Inc., Frisco, CO
Perry T. Jones (vehicle systems integration), Center for Transportation Analysis, Oak Ridge
National Labs, Knoxville, TN
Arun Kashyap (sustainable energy systems, climate change, CDM, private sector
involvement), United Nations Development Programme (UNDP), New York
Lawrence L. Kazmerski (solar, photovoltaic), Science and Technology Partnerships, National
Renewable Energy Laboratory (NREL), Golden, CO
Madhu Khanna (voluntary approaches to pollution control, welfare analysis, policies for
carbon sequestration), Department of Agricultural and Consumer Economics, Energy
Biosciences Institute, Institute of Genomic Biology, University of Illinois, Urbana, IL
Israel Koren (green computing), University of Massachusetts, Amherst, MA
Chung K. Law (hydrogen combustion), Department of Mechanical and Aerospace
Engineering, Princeton University, Princeton, NJ
Hongtan Liu (solar energy and hydrogen energy technology, fuel cells), Clean Energy
Research Institute, Department of Mechanical and Aerospace Engineering, University of
Miami, FL
Chang Mei (wave power), Department of Civil & Environmental Engineering, Massachusetts
Institute of Technology (MIT), Cambridge, MA
Editorial board xix

James Miller (advanced batteries, fuel cells, and hydrogen technologies for electric vehicles
and stationary applications), Office of Vehicle Technologies, United States Department of
Energy, Argonne National Laboratory, Argonne, IL
Daniel Mosse (green computing/sustainable computing), University of Pittsburgh,
Pittsburgh, PA
Emily Nelson (biofuels, green aviation, numerical modeling), Bio Science and Technology
Branch, NASA Glenn Research Center, Cleveland, OH
Kaushik Rajashekara (power electronics & drives and fuel cell power conversion), School of
Engineering & Computer Science, University of Texas at Dallas, Dallas, TX
Sanjay Ranka (green computing), University of Florida, Gainesville, FL
Athena Ronquillo- Ballesteros (international climate policy: climate finance, sustainable
energy and reform), World Resources Institute & Green Independent Power Producers
Network, Washington DC
Jack Rosebro (electric, hybrid plug-in, and hybrid vehicles), Los Angeles, CA
Subhash C. Singhal (fuel cells), Director, Fuel Cells, Energy and Environment Directorate,
Pacific Northwest National Laboratory, Richland, WA
Gregory Stephanopoulos (renewable fuels), W.H. Dow Professor of Chemical Engineering
and Biotechnology, Massachusetts Institute of Technology (MIT), Cambridge, MA
Lawrence E. Susskind (mediation of regulatory disputes, technology negotiation, policy
dialogue, stakeholder engagement), Department of Urban Studies and Planning,
Massachusetts Institute of Technology (MIT), Cambridge, MA
Veerle Vandeweerd (climate change), Director, Energy and Environment Group, United
Nations Development Programme (UNDP), New York
Peter F. Varadi (solar energy (PV)), P/V Enterprises, Inc., Chevy Chase, MD
YEMEN
Hussain Al-Towaie (solar power for seawater desalination), Owner & CEO at Engineering
Office “CE&SD” (Clean Environment & Sustainable Development), Aden
This page intentionally left blank
Table of contents

About the book series vii


Editorial board ix
Contributors xxix
Editors’ foreword xxxiii
About the editors xxxv
Acknowledgements xxxvii

1. Decontamination of water by solar irradiation 1


Sixto Malato, Manuel I. Maldonado, Pilar Fernández, Isabel Oller &
Inmaculada Polo
1.1 Introduction 1
1.2 Solar advanced oxidation processes 3
1.2.1 TiO2 solar photocatalysis 4
1.2.2 Solar photo-fenton 5
1.3 Solar technical issues 6
1.3.1 Hardware for solar AOPs 6
1.3.2 Solar photocatalytic treatment plants 10
1.4 Treatment of industrial wastewaters 11
1.4.1 Toxicity and biodegradability assessment 12
1.4.2 Industrial wastewater treatment by combined
AOPs/biotreatment 14
1.5 Treatment of secondary effluents 16
1.6 Conclusions 18

2. Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis:


An innovative way of arsenic removal 23
Marta I. Litter, Ivana K. Levy, Natalia Quici, Martín Mizrahi, Gustavo Ruano,
Guillermo Zampieri & Félix G. Requejo
2.1 Introduction 23
2.2 Experimental section 26
2.2.1 Materials and methods 26
2.2.2 Irradiation systems 27
2.3 Results 27
2.3.1 As(V) photocatalytic experiments 27
2.3.2 As(III) photocatalytic experiments 29
2.3.3 Analysis of solid residues 31
2.4 Discussion 34
2.4.1 Mechanisms at acid pH 34
2.4.2 Effect of pH 36
2.4.3 Comparison with previous results 38
2.5 Conclusions 38

xxi
xxii Table of contents

3. Synthesis, characterization and catalytic evaluation of tungstophosphoric acid


immobilized on Y zeolite 43
Candelaria Leal Marchena, Silvina Gomez, Liliana B. Pierella & Luis R. Pizzio
3.1 Introduction 43
3.2 Experimental 44
3.2.1 Samples preparation 44
3.2.2 Sample characterization 44
3.2.2.1 Textural properties 44
3.2.2.2 Nuclear magnetic resonance spectroscopy 45
3.2.2.3 Fourier transform infrared spectroscopy 45
3.2.2.4 X-Ray diffraction 45
3.2.2.5 Thermogravimetric analysis and differential scanning
calorimetry 45
3.2.2.6 Diffuse reflectance spectroscopy 45
3.2.2.7 Potentiometric titration 45
3.2.3 Photodegradation reaction 45
3.3 Results and discussion 46
3.4 Conclusions 54

4. Kinetic aspects of the photodegradation of phenolic and lactonic biocides


under natural and artificial conditions 59
Juan P. Escalada, Adriana Pajares, Mabel Bregliani, Alicia Biasutti,
Susana Criado, Patricia Molina, Walter Massad & Norman A. García
4.1 Introduction 59
4.2 Photochemical degradation 59
4.2.1 Modeling natural photodegradation 60
4.2.2 Artificial photodegradation 60
4.2.3 Biocides selected for the study 61
4.2.3.1 State of the art 62
4.3 Methods for photodegradation studies 63
4.3.1 Sensitized photoirradiation 64
4.3.1.1 Quenching of 1 Rf* and 3 Rf* 67
4.3.1.2 Quenching of O2 (1 g ) 71
4.3.2 Direct photolysis of ABA, BXN and DCP 75
4.4 Conclusions 76

5. Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst in a


recirculating batch reactor 81
Natalia Inchaurrondo, Josep Font & Patricia Haure
5.1 Introduction 81
5.2 Experimental 84
5.2.1 Catalyst preparation and characterization 84
5.2.2 Fenton like oxidation of phenol aqueous solutions 85
5.2.2.1 Reaction set-up 85
5.2.3 Analytical methods 86
5.3 Results and discussion 86
5.3.1 Blank experiment 86
5.3.2 Activity and stability tests 86
5.3.3 Deactivation phenomena 91
5.3.4 Effect of intermediate products adsorption 91
5.3.5 Initial pH effect 91
5.3.6 Copper load effect 92
5.3.7 Liquid flow rate effect 92
5.4 Conclusions 94
Table of contents xxiii

6. Degradation of a mixture of glyphosate and 2,4-D in water solution employing the


UV/H2 O2 process, including toxicity evaluation 99
Melisa Mariani, Roberto Romero, Alberto Cassano & Cristina Zalazar
6.1 Introduction 99
6.2 Materials and methods 100
6.2.1 Chemicals 100
6.2.2 Experimental setups and procedures 100
6.2.3 Analytical measurements 101
6.2.4 Toxicity assay 102
6.2.5 Operation 102
6.3 Results and discussion 102
6.3.1 Preliminary runs 102
6.3.2 Effect of initial pH values 103
6.3.3 Effects of initial hydrogen peroxide concentration 106
6.3.4 Effect of glyphosate and 2,4-D initial concentrations 107
6.3.5 Effect of variations in the incident UV spectral fluence rate at the
irradiated reactor walls 108
6.3.6 Total organic carbon (TOC) evolution 110
6.3.7 Formation of by-products and intermediates 110
6.3.8 Toxicity and chemical oxygen demand assays 112
6.4 Conclusions 114

7. Degradation of perchlorate dissolved in water by a combined application of


ion exchange resin and zerovalent iron nanoparticles 119
Luis Cumbal, Daniel Delgado & Erika Murgueitio
7.1 Introduction 119
7.2 Experimental section 120
7.2.1 Chemicals 120
7.2.2 Procedures 120
7.2.2.1 Preparation of nanoparticles 120
7.2.2.2 Physical characterization of nanoparticles 120
7.2.2.3 Conditioning of ion exchange resins and loading with
perchlorate 121
7.2.2.4 Kinetic tests 121
7.2.2.5 Degradation of perchlorate 121
7.2.3 Chemical analysis 122
7.3 Results and discussion 122
7.3.1 Physical characterization of nanoparticles 122
7.3.2 Kinetic tests 123
7.3.3 Degradation of perchlorate 124
7.3.4 Effect of competing ions and organic matter on the degradation of
perchlorate 125
7.4 Conclusions 127

8. Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 131
Pamela Yanina González Clar, Gustavo Levin, María Victoria Miranda &
Viviana Campo Dall’ Orto
8.1 Introduction 131
8.2 DB273 enzymatic decoloration 133
8.2.1 The enzyme 133
8.2.2 Color removal by oxidation 134
8.3 DB273 discoloration by adsorption 138
8.3.1 Synthesis and characterization of the polyampholyte 138
8.3.2 Kinetics of sorption 139
xxiv Table of contents

8.3.3 Isotherm data analysis 140


8.3.4 FTIR analysis 144
8.4 Conclusions 145

9. Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 149


Joana Femia, Melisa Mariani, Alberto Cassano, Cristina Zalazar & Inés Tiscornia
9.1 Introduction 149
9.2 Materials and methods 150
9.2.1 Chemicals 150
9.2.2 Experimental setups and procedures 151
9.2.3 Analytical methods 152
9.2.4 Bioassay test 153
9.3 Results and discussion 153
9.3.1 Preliminary runs 153
9.3.2 Effect of initial H2 O2 concentration 153
9.3.3 Total organic carbon (TOC) evolution 156
9.3.4 Evaluation of electrical energy per order 157
9.3.5 Toxicity evaluation 158
9.4 Conclusions 159

10. Abatement of nitrate in drinking water. A comparative study of photocatalytic


and conventional catalytic technologies 163
F. Albana Marchesini, Guadalupe Ortiz de la Plata, Orlando Alfano,
M. Alicia Ulla, Eduardo Miró & Alberto Cassano
10.1 Introduction 163
10.2 Materials and methods 165
10.2.1 Chemicals 165
10.2.2 Catalyst preparation 165
10.2.3 Catalyst characterization 165
10.2.3.1 X-Ray diffraction analysis (DRX) 165
10.2.3.2 Temperature-programmed reduction (TPR) 165
10.2.4 Catalytic activity measurements 165
10.2.4.1 Preliminary batch experiments 165
10.2.4.2 Photocatalytic experiments 166
10.2.5 Analytical methods 168
10.3 Results and discussion 168
10.3.1 Physicochemical characterization of the Pt,In/TiO2 catalyst 168
10.3.2 Catalytic reduction of nitrates: Conventional batch reactor 171
10.3.3 Catalytic reduction of nitrates: Photocatalytic reactor 172
10.3.4 Spatial distribution of the radiation absorption 174
10.4 Conclusions 175

11. Photocatalytic inactivation of airborne microorganisms. Performance of


different TiO2 coatings 179
Silvia Mercedes Zacarías, María Lucila Satuf, María Celia Vaccari &
Orlando Alfano
11.1 Introduction 179
11.2 Kinetic study 180
11.2.1 Experimental set up and procedure 180
11.2.2 Inactivation of spores 183
11.2.3 Kinetic modeling 183
11.2.3.1 Proposed kinetic model 183
11.2.3.2 Radiation model 185
Table of contents xxv

11.2.3.3 Kinetic parameters estimation 186


11.3 Study of different TiO2 coatings 187
11.3.1 Efficiency parameters 188
11.3.2 Preparation of the photocatalytic coatings 189
11.3.3 Characterization of the photocatalytic plates 189
11.3.4 Evaluation of photocatalytic efficiencies 190
11.3.5 Discussion 192
11.4 Conclusions 193

12. Water decontamination by heterogeneous photo-Fenton processes over iron,


iron minerals and iron-modified clays 197
Andrea De León, Marta Sergio, Juan Bussi, Guadalupe Ortiz de la Plata,
Alberto Cassano & Orlando Alfano
12.1 Introduction 197
12.2 Catalysts for use in heterogeneous photo-fenton processes 199
12.2.1 Iron and iron minerals 199
12.2.2 Supported and immobilized iron species 199
12.2.3 Iron species supported on clays 199
12.3 Experimental 200
12.3.1 Catalysts 200
12.3.1.1 Fe-PILCs 200
12.3.1.2 Goethite 200
12.3.1.3 Zerovalent iron 200
12.3.2 Catalyst characterization 200
12.3.3 Photocatalytic tests 201
12.3.3.1 Fluidized bed batch reactor 201
12.3.3.2 Stirred batch reactor 201
12.3.4 Analytical techniques 201
12.4 Catalytic activity 201
12.4.1 Iron-pillared clays used for dye degradation 201
12.4.1.1 Contribution of different processes entailed in
contaminant removal 201
12.4.1.2 Influence of the clay aggregate size used for Fe-PILC
preparation 203
12.4.1.3 Influence of the initial pH of the reaction medium 205
12.4.1.4 Selection of the temperature for calcination of the
exchanged clay 208
12.4.2 Fe-PILC, goethite and zerovalent iron in 2-chlorophenol
degradation 210
12.5 Conclusions 214

13. Modified montmorillonite in photo-Fenton and adsorption processes 217


Lucas M. Guz, Melisa Olivelli, Rosa M. Torres Sánchez,
Gustavo Curutchet & Roberto J. Candal
13.1 Introduction 217
13.2 Experimental section 219
13.2.1 Materials 219
13.2.2 Iron (III) modified montmorillonite (Fe-MMT) 220
13.2.3 Copper (II) modified montmorillonite (Cu-MMT) 220
13.2.4 Biomodified montmorillonite (Apha-BMMT) 220
13.2.5 Adsorption of Cu(II) on MMT and Apha-BMMT 220
13.2.6 Materials characterization 221
13.2.7 Photo-Fenton experiments 221
xxvi Table of contents

13.3 Results 221


13.3.1 Adsorption of Cu(II) on P5-MMT and Apha-BMMT 221
13.3.2 Catalysts characterization 222
13.3.3 Photo-Fenton experiments 224
13.4 Discussion 228
13.5 Conclusions 232

14. Photocatalytic degradation of dichlorvos solution using TiO2 -supported


ZSM-11 zeolite 235
Silvina Gomez, Candelaria Leal Marchena, Luis Pizzio & Liliana Pierella
14.1 Introduction 235
14.2 Experimental 236
14.2.1 Preparation of zeolite supported TiO2 catalyst 236
14.2.2 Characterization of the photocatalysts 236
14.2.3 Photocatalytic experiments and analyses 237
14.3 Results and discussion 237
14.3.1 Characterization of TiO2 /zeolite catalysts 237
14.3.1.1 XRD analysis 237
14.3.1.2 FTIR spectra 240
14.3.1.3 BET surface area 240
14.3.2 Photocatalytic evaluation 241
14.3.2.1 Preliminary studies 241
14.3.2.2 Effect of TiO2 content on TiO2 /HZSM-11 and
TiO2 /NH4 -ZSM-11 samples 242
14.3.2.3 Effect of the preparation of the catalyst and role of the
support 244
14.3.2.4 Effect of catalyst amount 245
14.3.2.5 Effect of the calcination temperature 245
14.3.2.6 Effect of initial pH value 246
14.3.2.7 Effect of adding H2 O2 to the photodegradation of DDVP 247
14.3.3 Photocatalyst recycling studies 248
14.4 Conclusions 250

15. Water disinfection with UVC and/or chemical inactivation. Mechanistic


differences, implications and consequences 253
Marina Flores, Rodolfo Brandi, Alberto Cassano & Marisol Labas
15.1 Introduction 253
15.2 Disinfection 254
15.3 UV disinfection 254
15.3.1 The principle of UV disinfection 255
15.3.1.1 Repair mechanisms 256
15.3.2 Case study: UV disinfection in clear water conditions 257
15.3.2.1 Experimental procedure 257
15.3.2.2 Experimental runs 257
15.3.2.3 Kinetic model 258
15.3.2.4 Experimental results 260
15.4 Hydrogen peroxide 261
15.4.1 The principle of disinfection using hydrogen peroxide 261
15.4.2 Case study: hydrogen peroxide disinfection in clear water conditions 261
15.4.2.1 Experimental procedure 262
15.4.2.2 Kinetic model 262
15.4.2.3 Mathematical model final equations 264
15.4.2.4 Experimental results 264
Table of contents xxvii

15.5 Peracetic acid 265


15.5.1 PAA mode of action 266
15.5.2 Case study: water disinfection with peracetic acid in clear water
conditions 266
15.5.2.1 Experimental procedure 266
15.5.2.2 A proposed kinetics of peracetic acid decomposition 267
15.5.2.3 Experimental results 267
15.6 Peracetic acid + UV light 268
15.6.1 Case study: disinfection of water with peracetic acid and its
combination with UVC 268
15.6.1.1 Experimental procedure 268
15.6.1.2 A proposed kinetics of peracetic acid + UV 269
15.6.1.3 Experimental results 269
15.7 Hydrogen peroxide + UV 270
15.7.1 Case study: disinfection with hydrogen peroxide and UV light in
clear water conditions 270
15.7.1.1 Experimental procedure 270
15.7.1.2 Kinetic model 271
15.7.1.3 Experimental results 271
15.8 Conclusions 272
Appendix 273

16. Ag/AgCl composite material: synthesis, characterization and application


in treating wastewater 281
Wei-Lin Dai, Quan-Jing Zhu, Jian-Feng Guo & Bo-Wen Ma
16.1 Introduction 281
16.2 Synthesis of the photocatalysts 282
16.2.1 Ag/AgCl core-shell sphere 282
16.2.1.1 Preparation of Ag spheres using ascorbic acid as
the reducing agent 282
16.2.1.2 Preparation of Ag/AgCl core-shell sphere using ferric
chloride 283
16.2.2 Ag/AgCl@Cotton-fabric 283
16.2.3 Ag-AgCl/WO3 hollow sphere 283
16.2.3.1 Preparation of the hollow sphere PbWO4 283
16.2.3.2 Preparation of the hollow sphere WO3 283
16.2.3.3 Preparation of Ag-AgCl/WO3 283
16.2.4 Ag-AgCl@TiO2 284
16.2.5 Ag-AgI/Fe3 O4 @SiO2 284
16.2.5.1 Synthesis of Fe3 O4 particles 284
16.2.5.2 Synthesis of Fe3 O4 @SiO2 microspheres 284
16.2.5.3 Synthesis of AgI/Fe3 O4 @SiO2 284
16.2.5.4 Synthesis of Ag-AgI/Fe3 O4 @SiO2 284
16.3 Characterization of the photocatalysts 285
16.4 Evaluation of photocatalytic activity 285
16.5 Results and discussion 285
16.5.1 Ag/AgCl core-shell sphere 285
16.5.2 Ag/AgCl@Cotton-fabric 291
16.5.3 Ag-AgCl/WO3 hollow sphere 293
16.5.4 Ag-AgCl@TiO2 295
16.5.5 Ag-AgI/Fe3 O4 @SiO2 297
16.6 Conclusions 299
xxviii Table of contents

17. Highly photoactive Er3+ -TiO2 system by means of up-conversion and


electronic cooperative mechanism 303
Sergio Obregón & Gerardo Colón
17.1 Introduction 303
17.2 Experimental section 303
17.2.1 Synthesis of photocatalysts 303
17.2.2 Materials characterization 304
17.2.3 Photocatalytic experimental details 304
17.3 Results and discussion 305
17.4 Conclusions 311

18. Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 313
Elias Stathatos, Dimitrios Papoulis & Dionisios Panagiotaras
18.1 Introduction 313
18.2 TiO2 nanoparticles and films 314
18.2.1 Sol-gel method for nanoparticles and films 314
18.2.2 Hydrothermal route for TiO2 nanoparticles and films 315
18.3 Stabilized TiO2 particles with sol-gel method on clay minerals. Palygorskite
clay mineral as support for TiO2 particles 316
18.3.1 Materials and methods 316
18.3.2 Photocatalyst characterization 317
18.3.3 Photocatalytic activity of sol-gel TiO2 modified palygorskite clay
mineral for polluted water with an azo dye 319
18.4 Stabilized TiO2 particles with hydrothermal route on clay minerals.
Halloysite clay mineral as an example 320
18.4.1 Materials and methods 320
18.4.2 Photocatalyst characterization 322
18.4.3 Photocatalytic activity of TiO2 modified halloysite clay mineral
for air purification 322
18.5 Conclusions 324

19. Photodegradation of beta-blockers in water 327


Virender K. Sharma, Hyunook Kim & Radek Zboril
19.1 Introduction 327
19.2 Phototransformation in water 328
19.3 Influence of water chemistry 330
19.3.1 pH 330
19.3.2 Nitrate ion 330
19.3.3 Types of natural organic matter 331
19.4 Mechanism 333
19.5 Mineralization and toxicity 333
19.6 Conclusions 334

20. Final conclusions 337


Marta I. Litter, Roberto J. Candal & J. Martín Meichtry

Subject index 341

Book series page 349


Contributors

Orlando Alfano: National Institute of Chemical Technology (INTEC), National University of


Litoral-CONICET, Santa Fe, Argentina & Faculty of Engineering and Hydrical Sciences, National
University of Litoral, Santa Fe, Argentina
Alicia Biasutti: Chemistry Department, National University of Río Cuarto, Córdoba, Argentina
Rodolfo Brandi: National Institute of Catalysis and Petrochemistry (INCAPE), National Univer-
sity of Litoral-CONICET, Santa Fe, Argentina
Mabel Bregliani: Río Gallegos Academic Unit, National University of Austral Patagonia, Santa
Cruz, Argentina
Juan Bussi: Laboratory of Physicochemistry of Surfaces, DETEMA, Faculty of Chemistry,
University of the Republic, Montevideo, Uruguay
Roberto J. Candal: School of Science and Technology, National University of General San Martín,
Prov. of Buenos Aires & Institute of Physical Chemistry of Materials, Environment and Energy
(INQUIMAE), Faculty of Exact and Natural Sciences, University of Buenos Aires-CONICET,
Buenos Aires, Argentina
Alberto Cassano: National Institute of Chemical Technology (INTEC), National University of
Litoral-CONICET, Santa Fe, Argentina & Department of Environment, Faculty of Engineering
and Hydrical Sciences, National University of Litoral, Santa Fe, Argentina
Gerardo Colón: Institute of Materials Science of Sevilla, Centro Mixto CSIC-University of Sevilla,
Sevilla, Spain
Susana Criado: Department of Chemistry, National University of Río Cuarto, Córdoba, Argentina
Luis Cumbal: Center of Scientific Research, Army Polytechnical School (ESPE), Sangolqui,
Ecuador
Gustavo Curutchet: School of Science andTechnology, National University of General San Martín,
Prov. of Buenos Aires & CONICET, Buenos Aires, Argentina
Wei-Lin Dai: Department of Chemistry, Fudan University, Shanghai, P.R. China
Daniel Delgado: Center of Scientific Research, Army Polytechnical School (ESPE), Sangolqui,
Ecuador
Viviana Campo Dall’ Orto: Department of Analytical Chemistry and Physical Chemistry, Faculty
of Pharmacy and Biochemistry, University of BuenosAires & CONICET, BuenosAires, Argentina
Andrea De León: Laboratory of Physicochemistry of Surfaces, DETEMA, Faculty of Chemistry,
University of the Republic, Montevideo, Uruguay
Juan P. Escalada: Río Gallegos Academic Unit, National University of Austral Patagonia, Santa
Cruz, Argentina
Joana Femia: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET, Santa Fe, Argentina
xxix
xxx Contributors

Pilar Fernández: Solar Platform of Almería, Center of Energetic, Environmental and Technolog-
ical Research (CIEMAT), Almería, Spain
Marina Flores: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET, Santa Fe, Argentina
Josep Font: Department of Chemical Engineering, Universitat Rovira i Virgili, Tarragona, Spain
Norman A. García: Department of Chemistry, National University of Río Cuarto, Córdoba,
Argentina
Silvina Gomez: Zeolites Group, Center of Chemical Research and Technology (CITeQ), National
Technological University, Córdoba Regional University, Córdoba, Argentina & CONICET,
Buenos Aires, Argentina
Pamela Yanina González Clar: Department of Analytical Chemistry and Physical Chemistry,
University of Buenos Aires, Buenos Aires, Argentina
Jian-Feng Guo: Department of Chemistry, Fudan University, Shanghai, P.R. China
Lucas M. Guz: School of Science and Technology, National University of General San Martín,
Prov. of Buenos Aires & CONICET, Buenos Aires, Argentina
Natalia Inchaurrondo: Catalysts and Surfaces Division – Institute of Technology of Materials
(INTEMA)-CONICET, National University of Mar del Plata, Buenos Aires, Argentina
Patricia Haure: Catalysts and Surfaces Division – Institute of Technology of Materials (INTEMA)-
CONICET, National University of Mar del Plata, Buenos Aires, Argentina
Hyunook Kim: Department of Environmental Engineering, The University of Seoul, Seoul, Korea
Marisol Labas: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET, Santa Fe, Argentina
Candelaria Leal Marchena: Zeolites Group, Center of Chemical Research and Tecnology
(CITeQ), National Technological University, Córdoba Regional University, Córdoba, Argentina &
CONICET, Buenos Aires, Argentina
Gustavo Levin: Department of Microbiology, Inmunology and Biotechnology, Faculty of Phar-
macy and Biochemistry, University of Buenos Aires & CONICET, Buenos Aires, Argentina
Ivana K. Levy: Remediation Technologies Division, Environmental Chemistry Department,
Chemistry Management, NationalAtomic Energy Commission, BuenosAires; National Scientific
and Technique Research Council (CONICET), Argentina
Marta I. Litter: Remediation Technologies Division, Environmental Chemistry Department,
Chemistry Management, National Atomic Energy Commission, Buenos Aires; National Sci-
entific and Technique Research Council (CONICET); Institute of Research and Environmental
Engineering, National University of General San Martín, Argentina
Bo-Wen Ma: Department of Chemistry, Fudan University, Shanghai, P.R. China
Sixto Malato: Solar Platform of Almería, Center of Energetic, Environmental and Technological
Research (CIEMAT), Almería, Spain
Manuel I. Maldonado: Solar Platform of Almería, Center of Energetic, Environmental and
Technological Research (CIEMAT), Almería, Spain
F. Albana Marchesini: National Institute of Catalysis and Petrochemistry (INCAPE), National
University of Litoral-CONICET & Faculty of Chemical Engineering, National University of
Litoral, Santa Fe, Argentina Santa Fe, Argentina
Melisa Mariani: National Institute of Catalysis and Petrochemistry (INCAPE), National Univer-
sity of Litoral–CONICET, Santa Fe, Argentina
Contributors xxxi

Walter Massad: Departament of Chemistry, National University of Río Cuarto, Córdoba,


Argentina
María Victoria Miranda: Department of Microbiology, Immunology and Biotechnology, Faculty
of Pharmacy and Biochemistry, University of BuenosAires & CONICET, BuenosAires, Argentina
Eduardo Miró: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET & Faculty of Chemical Engineering, National University of Litoral, Santa
Fe, Argentina Santa Fe, Argentina
Martín Mizrahi: Institute of Theoretical and Applied Physicochemical Research, La Plata,
CONICET, Buenos Aires & Faculty of Exact Sciences, National University of La Plata, La
Plata, Argentina
Patricia Molina: Departament of Chemistry, National University of Río Cuarto, Córdoba,
Argentina
Erika Murgueitio: Center of Scientific Research, Army Polytechnical School (ESPE), Sangolqui,
Ecuador
Sergio Obregón: Institute of Materials Science of Sevilla, Centro Mixto CSIC-University of
Sevilla, Sevilla, Spain
Melisa Olivelli: Center of Technology of Mining Resources and Ceramics (CETMIC),
CIC-CONICET, M. B. Gonnet, Prov. Buenos Aires, Argentina & School of Science and
Technology, National University of General San Martín, Prov. of Buenos Aires & CONICET,
Buenos Aires, Argentina
Isabel Oller: Solar Platform of Almería, Center of Energetic, Environmental and Technological
Research (CIEMAT), Almería, Spain
Guadalupe Ortiz de la Plata: National Institute of Chemical Technology (INTEC), National
University of Litoral-CONICET, Santa Fe, Argentina
Adriana Pajares: Río Gallegos Academic Unit, National University of Austral Patagonia, Santa
Cruz, Argentina & Faculty of Engineering, National University of Patagonia San Juan Bosco,
Chubut, Argentina
Dionisios Panagiotaras: Mechanical Engineering Department, Technological-Educational Insti-
tute of Patras, Patras, Greece
Dimitrios Papoulis: Department of Geology, University of Patras, Patras, Greece
Liliana B. Pierella: Zeolites Group, Center of Chemical Research and Tecnology (CITeQ),
National Technological University, Córdoba Regional University, Córdoba, Argentina &
CONICET, Buenos Aires, Argentina
Luis R. Pizzio: Center of Research and Development in Applied Sciences “Dr. J.J. Ronco”
(CINDECA), National University of La Plata, La Plata Argentina & CONICET, Buenos Aires,
Argentina
Inmaculada Polo: Solar Platform of Almería, Center of Energetic, Environmental and Technolog-
ical Research (CIEMAT), Almería, Spain
Natalia Quici: Remediation Technologies Division, Environmental Chemistry Department,
Chemistry Management, NationalAtomic Energy Commission, BuenosAires; National Scientific
and Technique Research Council (CONICET), Buenos Aires, Argentina
Félix G. Requejo: Institute of Theoretical and Applied Physicochemical Research, La Plata,
CONICET, Buenos Aires & Faculty of Exact Sciences, National University of La Plata, La Plata,
Argentina
xxxii Contributors

Roberto Romero: National Institute of Catalysis and Petrochemistry (INCAPE), National


University of Litoral-CONICET, Santa Fe, Argentina
Gustavo Ruano: Bariloche Atomic Center and Balseiro Institute, Bariloche, Río Negro Prov.,
Argentina
María Lucila Satuf: National Institute of Chemical Technology (INTEC), National University of
Litoral-CONICET, Santa Fe, Argentina
Marta Sergio: Laboratory of Physicochemistry of Surfaces, DETEMA, Faculty of Chemistry,
University of the Republic, Montevideo, Uruguay
Virender K. Sharma: Chemistry Department, Florida Institute of Technology, Melbourne,
Florida, USA
Elias Stathatos: Electrical Engineering Department, Technological-Educational Institute of Patras,
Patras, Greece
Inés Tiscornia: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET, Santa Fe, Argentina
Rosa M. Torres Sánchez: Centro de Tecnología de Recursos Mineros y Cerámica (CETMIC),
CIC-CONICET, M. B. Gonnet, Prov. de Buenos Aires, Argentina
M. Alicia Ulla: National Institute of Catalysis and Petrochemistry (INCAPE), National University
of Litoral-CONICET, and Chemical Engineering Faculty, National University of Litoral, Santa
Fe, Argentina
María Celia Vaccari: Faculty of Biochemistry and Biological Sciences, National University of
Litoral, Santa Fe, Argentina
Silvia Mercedes Zacarías: National Institute of Chemical Technology (INTEC), National Univer-
sity of Litoral-CONICET, Santa Fe, Argentina & Faculty of Biochemistry and Biological Sciences,
National University of Litoral, Santa Fe, Argentina
Cristina Zalazar: National Institute of Chemical Technology (INTEC), National University of
Litoral-CONICET, Santa Fe, Argentina & Department of Environment, Faculty of Engineering
and Hydrological Sciences, National University of Litoral, Santa Fe, Argentina
Guillermo Zampieri: Bariloche Atomic Center & Instituto Balseiro, Bariloche, Prov. de Río
Negro, and CONICET, Buenos Aires Argentina
Radek Zboril: Regional Centre of Advanced Technologies and Materials, Departments of Physical
Chemistry and Experimental Physics, Faculty of Science, Palacky University, Olomouc, Czech
Republic
Quan-Jing Zhu: Department of Chemistry, Fudan University, Shanghai, P.R. China
Editors’ foreword

This book gives a state-of-the-art overview on environmental applications of advanced oxidation


technologies (AOTs) as sustainable, low cost and low energy consuming treatments for water, air,
and soil. It includes information on innovative research and development on TiO2 photocatalytic
redox processes, Fenton, Photo-Fenton processes, zerovalent iron technology, etc., highlighting
possible applications of AOTs in developing and industrialized countries around the world in the
framework of “A crosscutting and comprehensive look at environmental problems”.
Examples of non-conventional, advanced oxidation technologies (AOTs) and advanced reduc-
tion technologies (ARTs), as sustainable solutions for environmental treatments are presented,
selected from recognized groups of different countries. AOTs are well known for their capacity
for oxidizing and mineralizing almost any organic contaminant and transforming several inor-
ganic pollutants strongly resistant to conventional treatments, such as metals and metalloids. The
AOTs presented in this book are mainly TiO2 photocatalysis (HP) and photo-Fenton, but others
like UV/H2 O2 , enzyme/H2 O2 , peracetic acid and UV photolysis are also exemplified. ARTs
are emerging technologies from which reductive TiO2 heterogeneous photocatalysis and use of
zerovalent iron nanoparticles are increasingly studied in last times. The use of zerovalent iron,
especially in the form of nanoparticles, is a very economic novel tool for removal of contam-
inants from water and has been successfully tested in batch and in-situ treatments, while TiO2
heterogeneous photocatalysis is especially investigated for the treatment of organic pollutants and
heavy metal ions.
Along the different chapters of this book an overview of the latest developments in AOTs
and ARTs and the trends followed by different research groups to improve the efficiency of the
processes are discussed. AOTs and ARTs are not universal solutions for water and gas treatments,
but they represent an alternative to completely eliminate or increase the biodegradability of
several recalcitrant compounds. These technologies are especially attractive in countries with
high solar irradiation where photo-based technologies can be used, diminishing the energy cost.
Latin-American, African, South-European and several Asian countries, having serious pollution
problems, can take benefits from these technologies.
Chapter 1 is a review showing the different approaches based on the use of sunlight as eco-
nomical and sustainable irradiation source while Chapter 4 is another review article presenting a
comparative kinetic study between naturally and artificially promoted photodegradation of phe-
nolic and lactonic biocides, which can be decomposed either by direct or by indirect photolysis.
In line with these studies, Chapter 19 describes the photodegradation of pharmaceuticals such as
β-blockers using simulated sunlight in the presence and absence of photosensitizers.
Chapter 7 shows the combined use of an ion exchange resin (A530E) and iron nanosized
particles for removal of perchlorate from water. Chapter 2 describes the reduction of As(V) and
As(III) species by TiO2 photocatalysis under anoxic conditions as an innovative way for arsenic
removal from water. Chapter 10 discusses the abatement of nitrate in drinking water through a
comparative study of photocatalytic and conventional catalytic technologies.
Chapters 3, 14, 16, 17 and 18 describe the preparation and characterization of modified TiO2
materials or other catalysts useful for AOTs/ARTs. The emphasis was put on increasing the
photocatalytic activity and on shifting the light absorption edge to higher wavelengths.
Other chapters refer to Fenton or photo-Fenton as reactions that use heterogeneous catalysts
containing iron or copper supported on different substrates. These different approaches were
xxxiii
xxxiv Editors’ foreword

explored in order to improve mineralization, increase the operation pH and facilitate catalyst
separation.
The combination of UVC irradiation and H2 O2 is a well-known process for water purification
that does not need catalysts and may work at different pH values. Chapters 6 and 9 exemplify
the application of this process to the elimination of pesticides in water, trying to find the best
conditions to treat solutions containing typical concentrations of the herbicides glyphosate and
2,4-D and chlorpyrifos, respectively.
Decomposition of H2 O2 to generate HO• can also be catalyzed by enzymes. Chapter 8
describes the discoloration of Direct Blue 273 using enzymatic discoloration and adsorption.
Both technologies proved to be very effective and economic.
Water disinfection is of great concern in the entire world, and large efforts are made in the
investigation of alternative disinfection/oxidation methods. Chapter 11 describes the reduction of
the viability of Bacillus subtilis spores over UV-irradiated TiO2 films in the gas phase.
Chapter 15 is a brief account of the possibilities of peracetic acid and hydrogen peroxide as
additional optional choices. Five disinfection methods were compared: disinfection methods: (i)
UV disinfection, (ii) H2 O2 disinfection, (iii) peracetic acid disinfection (iv) peracetic acid + UV
disinfection and (v) H2 O2 + UV disinfection. The preliminary results suggest that peracetic acid
may be considered as an environmentally friendly good alternative for water treatment.
The book is aimed to professionals and academics worldwide, working in the areas of water
resources, water supply, environmental protection, etc. It is a useful information source for deci-
sion and policy makers, and other stakeholders working on solutions for environmental problems.

Marta I. Litter
Roberto J. Candal
J. Martín Meichtry
January 2014
About the editors

Professor Marta I. Litter was born in BuenosAires, Argentina. She holds a degree and a Doctorate
in Chemistry from the University of Buenos Aires, Argentina. She performed a Postdoctoral stage
at the University of Arizona, USA, in Polymer Chemistry (1983). She is the Head of Remediation
Technologies Division, National Atomic Energy Commission, Argentina, Principal Researcher,
National Research Council (CONICET, Argentina) and Full Professor of the University of General
San Martín, Argentina. She has written more than 150 scientific publications. She has coordinated
several projects on water treatment, mainly in Advanced Oxidation Technologies. She was also
Coordinator of the CYTED IBEROARSEN Network (2006–2009).
She received the Mercosur Prize 2006 in Science and Technology, Technologies for Social
Inclusion, for the Project: “Potabilization of water by low-cost technologies in isolated rural
zones of Mercosur” and the Mercosur Prize 2011 in Science and Technology, Technologies for
Sustainable Development, for the Project: “The problem of arsenic in the Mercosur. An integrated
and multidisciplinary approach to contribute to its resolution.”
At present she is President of the Local Organizing Committee of the 5th International Congress
on Arsenic in the Environment (As2014) to be held in Buenos Aires, Argentina, from 11 to 16
May 2014.

xxxv
xxxvi About the editors

Professor Roberto J. Candal, born 1960 in Argentina, holds a degree in Chemistry and a
doctorate in the field of Inorganic Chemistry from the University of Buenos Aires. He held
a three year position as Post-doc at the Water Chemistry Laboratory, University of Wisconsin,
Madison WI, USA. His interests in research are focused on the development of new materials with
application in water or air remediation, photocatalysis, sol-gel chemistry and water chemistry.
Dr. Candal is co-author of more than 50 scientific publications in peer reviewed international
journals and books. He has directed or co-directed three PhD Thesis; at present, he is directing
three PhD Thesis in environmental chemistry. Since 2010 he is Associate Professor at the National
University of San Martín, Argentina, and Independent Researcher at the National Research Coun-
cil of Argentina (INQUIMAE-CONICET). Dr. Candal is a founding member of Argentina Society
for Science and Environmental Technology (SACyTA).

Dr. J. Martín Meichtry was born in 1977 in Colón, Entre Ríos, Argentina. He is Doctor in
Engineering from the University of Buenos Aires (2011). Presently he is Researcher at the Reme-
diation Technologies Division, Chemistry Management, National Atomic Energy Commission,
Argentina, Assistant Researcher of the National Research Council of Argentina (CONICET) and
Assistant Professor at the Chemistry Department, Buenos Aires School of the National Techno-
logical University, Argentina. He is author of 10 scientific publications, mainly in international
journals of high impact in physical chemistry and environmental sciences, 4 chapters of books
and many technical reports. He has more than 50 presentations in national and international
congresses and other scientific meetings. He has participated in three prized presentations: Envi-
ronmental Chemistry session, VI Congress Latin America SETAC (2003), Innovar Prize from
MINCYT Argentina (2009) and Environmental Technology and Engineering section, COPIME
Environmental Science Congress (2011). He has participated in 16 projects on water treatment,
especially on Advanced Oxidation Technologies and more especially on Heterogeneous Photo-
catalysis. He is reviewer of the Chemical Engineering Science, Water Research and Chemical
Engineering Science (ELS).
Acknowledgements

The editors thank the National Scientific and Technical Research Council of Argentina
(CONICET), the National Agency for the Promotion of Science and Technology (ANPCyT),
the National Atomic Energy Commission of Argentina (CNEA) and the University of General
San Martin of Argentina (UNSAM). The editors and authors thank also the technical people of
the Taylor & Francis Group for their cooperation and the excellent typesetting of the manuscript.

xxxvii
This page intentionally left blank
CHAPTER 1

Decontamination of water by solar irradiation

Sixto Malato, Manuel I. Maldonado, Pilar Fernández, Isabel Oller & Inmaculada Polo

1.1 INTRODUCTION

Humankind has been changing, contaminating, and polluting their environment since the Pale-
olithic era. This first type of pollution was air pollution through the use of fire (Spengler and
Sexton, 1983). Later, in the Bronze Age and Iron Age, the forging of metals produced significant
air and soil pollution as can be seen in glacial core samples in Greenland (Honget al., 1996).
As early as the 9th century, Arab and Persian scientists have written about pollution and waste
handlings (Gari, 2002). The industrial revolution in the 19th century led to the environmental
pollution, as we know it today with its new industries and the enormous consumption of coal and
oil, which lead to air pollution and the discharge of chemicals and industrial wastes into waterways
(streams and rivers). After WWII, the matter of pollution came into public focus not only through
the matter of atomic fallout (atomic warfare and testing), but as well as through the great smog
event in 1952 in London, which killed between 4000 and 12,000 people (De Angelo and Black,
2008). Several environmental disasters happened in 20th century such as the Mercury poison-
ing of Minamata Bay (Japan) 1956, the Dioxin disaster in Seveso (Italy) 1976, the Love Canal
chemical waste dump (USA) 1978, Three Mile Island core meltdown (USA) 1979, the Union
Carbide gas leak in Bhopal (India) 1984, Thermonuclear Meltdown in Chernobyl (Ukraine) 1986
and Fukushima (Japan) 2011, numerous events of oil leaks and spills (Torrey Canyon 1967, Piper
Alpha 1988, Exxon Valdez 1989, the Gulf War 1991, Deepwater Horizon 2010, etc.). It was also
very important for the environment the use of more than 75000 m3 of Agent Orange in Vietnam
during 1962–1971 and the destructive impact of acid rain on limestone, plants and lakes, which
was discovered as early as 1852 by Robert Angus Smith (Seinfeld, 1998), but was not studied
widely until the late 1960s (Likens et al., 1996).
All these events raised public awareness on pollution and resulted in various laws and legisla-
tions to prevent future disasters and clean up polluted sites. In contrast to those big environmental
disasters, the impact of continuous discharge of industrial wastes, pesticides, fertilizers and food
industry pharmaceuticals were considered to be nearly nonexistent or without any possible neg-
ative consequences, due to the wide spread thought “the solution to pollution is dilution”, and
no thought was lost on the impact of unregulated substances like hormones, pharmaceuticals for
human use, etc. However, as early as the 1940s, there was awareness in the scientific commu-
nity that certain chemicals can act as endocrine disrupting substances, or can mimic hormones
(Schueler, 1946; Sluczewski and Roth, 1948), and the first report of pharmaceuticals in wastewa-
ter treatment plant effluents was published in 1977 (Hignite and Azarnoff, 1977). During the last
30 or so years, the analytical and environmental chemistry has focused more or less exclusively
on the “conventional contaminants” (CC), mainly pesticides, industrial chemicals, heavy metals,
etc. Although these contaminants pose a great threat to the environment when released, they
only represent a small percentage of chemicals that can be detected nowadays in the environment
(Daughton and Ternes, 1999). The types of potentially hazardous substances and their range of
concentration in the environment may vary considerably.
Conventional wastewater treatment plants, which make use of activated sludge systems to
reduce biodegradable chemical oxygen demand (COD), are not effective when industrial wastew-
aters have to be treated due to the high COD load and the presence of recalcitrant compounds. This
1
2 S. Malato et al.

drawback becomes particularly critical when a significant concentration of non-biodegradable


toxic compounds is also present, rendering the wastewater itself toxic to the microorganisms
responsible for the biodegradation. In such instances, environmental regulatory requirements can
be achieved by chemical oxidation pre- or post-treatments.
Post-treatments can be used as polishing step after a biological treatment in the case of wastew-
aters containing large amounts of biodegradable organics and small concentrations of recalcitrant
compounds (Oller et al., 2011). Pretreatments can be advisable in the opposite case, i.e. when
the amount of bioresistant toxic contaminants is greater than that of biodegradable matter. Once
biodegradability has been achieved, effluents can be transferred to a cheaper biological treatment.
The key is to minimize residence time and reagent consumption during the expensive chemical
oxidation stage by applying an optimized coupling strategy (Esplugas and Ollis, 1997; Sarria
et al., 2003).
The most widely used advanced oxidation treatments are commonly referred to as advanced oxi-
dation processes (AOPs). AOPs are generally defined as oxidation processes generating hydroxyl
radicals which, in turn, are responsible for organic degradation due to their strong oxidant power
(HO• + H+ + e− → H2 O; E◦ = 2.33 V). Most of the systems classified as AOPs make use of a
combination of either (I) two oxidants (O3 /H2 O2 ), or (II) a catalyst and an oxidant (Fe2+ /H2 O2 ),
(III) an oxidant and irradiation (UV/O3 , UV/H2 O2 ), (IV) irradiation and a catalyst (UV/TiO2 ),
(V) irradiation, a catalyst and an oxidant (UV/Fe2+ / H2 O2 ), or finally (VI) an oxidant (H2 O2 ) and
ultrasonic irradiation (Gogate and Pandit, 2004).The common drawback of such systems is the
high demand of electrical energy for devices such as ozonizers, UV lamps, ultrasound generators,
etc., and this makes such treatments economically disadvantageous. This is why, although AOPs
are well known for their capacity for oxidizing and mineralizing almost any organic contaminant,
commercial applications are still scarce.
Several promising cost-cutting approaches have been proposed based on a real integration of
the AOP as part of a whole treatment train or process. Other proposed cost-cutting measures are
the use of renewable energy sources, i.e. sunlight as irradiation source for TiO2 photocatalysis
and photo-Fenton, further advances and development of applied reactors have improved plant
operation and control strategies to raise the degree of automation and lower the operational costs
(Malato et al., 2009).Therefore, low cost AOP systems suitable to be combined with biologi-
cal processes are innovative options in this area. Investigation concerning such options would
undoubtedly contribute to enhance the applicability of combined AOP+Biological processes for
treating industrial wastewater. The solar approach is a logical consequence for AOP cost saving
to be applied especially in the Southern Europe regions.
The development of new analytical techniques like gas chromatography coupled with mass
spectrometry (GC-MS and GC-MS/MS), liquid chromatography coupled with mass spectrometry
(LC-MS, LC-TOF/MS, LC-MS/MS), which push the limit of quantification/detection further into
the low nano-gram range, permitted the detection and analysis of new organic substances and
its metabolites in environmental samples, in extremely low concentrations (Gómez et al., 2009;
Hogenboomet al., 2009; Petrovic and Barceló, 2006; Pietrogrande and Basaglia, 2007; Robles-
Molina et al., 2010).The so called “emerging contaminants” (ECs) are defined as a group of
organic substances which are not subject to restrictions of any kind, but may be candidates for
future regulations, depending on the investigative results on the effects on human health, aquatic
life forms and their presence in the environment. A wide range of compounds are considered to be
relevant emerging compounds, such as: detergents, pharmaceutical products and its metabolites,
personal care products, flame retardants, antiseptics, fragrances, industrial additives, steroids
and hormones, among others. The principal characteristic of these contaminants is that they do
not have to be persistent in the environment to cause negative effects on life forms, as their
possible high degradation and/or elimination are compensated through the constant release into
the environment.
For the majority of these “ECs” there is no available data on the ecological impact, risk assess-
ment and eco toxicological behavior. Therefore, it is difficult if not impossible to predict which
effects these contaminants will have on the human health and on aquatic organisms. Many of them
Decontamination of water by solar irradiation 3

Figure 1.1. Contaminants enter the environment through different routes.

are omnipresent and persistent, and they have been detected in effluents of municipal wastewa-
ter treatment plants, rivers, aquifers and even in drinking water (Martínez Bueno et al., 2012).
Contaminants enter the environment through different routes as can be seen in Figure 1.1. The
two main ones are untreated urban wastewaters and MWTP effluents, as most of the MWTP are
not designed to treat such compounds. Another big contributor is the agricultural sector with its
use of veterinary drugs and pesticides. These substances are released directly into surface waters,
and may even seep through the soil to contaminate groundwater. Another way of entering the
environment is through the application of sewage sludge onto fields and subsequent leaching and
runoffs into surface- and groundwater, as well as the leaching of landfills into the aquifer.
This chapter will overview not only the main solar AOPs (TiO2 photocatalysis and photo-
Fenton) but also their application in the treatment of industrial wastewaters containing
conventional contaminants (CC) and effluents from MWTP containing micro-pollutants and ECs.

1.2 SOLAR ADVANCED OXIDATION PROCESSES

Rate constants (kOH , r = kOH [HO• ] C) for most reactions involving hydroxyl radicals in aqueous
solution are usually on the order of 106 to 109 M−1 s−1 . They are also characterized by their
not-selective attack, which is a useful attribute for wastewater treatment and solution of pollution
problems. The versatility of AOPs is also enhanced by the fact that there are different ways of
producing hydroxyl radicals, facilitating compliance with the specific treatment requirements.
During an ideal treatment by solar AOPs, contaminants are structurally altered forming smaller
and more biodegradable compounds until complete mineralization is achieved. For the application
of AOPs in wastewater treatment, the objectives can be either to: (i) eliminate biorecalcitrant and
toxic compounds, (ii) increase the biodegradability of wastewater before applying conventional
biological process, (iii) reduce the level of toxicity and micropollutants in the effluent or (iv)
disinfect the wastewater instead of applying traditional disinfection methods such as chlorination,
which is known to generate carcinogenic and mutagenic by-products such as halomethanes (Rizzo,
2011). Heterogeneous photocatalysis and homogeneous photo-Fenton are based on the use of a
wide-band-gap semiconductor and addition of H2 O2 to dissolved iron salts, respectively, and
irradiation with UV-VIS light. Both processes are of special interest since sunlight can be used
4 S. Malato et al.

Figure 1.2. Publication treating photocatalysis and the share treating solar-driven photocatalysis (source:
www.scopus.com, September 2012 search terms “photocatalysis” (red) and “solar” (blue)
within these results).

for them. The publications regarding the photocatalytic process rose continuously over the last
years, surpassing meanwhile a total number of more than 4500 peer-reviewed publications per year
in 2011. Though such a simple search does not necessarily include every single article correctly, it
still serves to prove the general trend of an increasing interest of the scientific community. Figure
1.2 shows the evolution of these publication activities. This figure also illustrates that much of
the literature takes into account the possibility of driving the process with solar radiation. This
fact is due to the fact that a prior photocatalytic process seems to be the most apt of all AOPs to
be driven by sunlight.

1.2.1 TiO2 solar photocatalysis


Photocatalysis is defined as a “change in the rate of a chemical reaction or its initiation under
the action of UV, visible or infrared radiation in the presence of a substance, the photocata-
lysts that absorb light quanta and are involved in the chemical transformation of the reaction
partners” (Braslavsky et al., 2007). In 1972, Fujishima and Honda discovered the possibility of
water splitting by means of a photoelectrochemical cell consisting of a TiO2 photoanode and a
Pt counter electrode (Fujishima and Honda, 1972). Water purification by means of illuminating
TiO2 was first proposed by Frank and Bard for cyanide and sulfite removal (Frank and Bard,
1977). Applications of heterogeneous photocatalysis include air purification, self-cleaning sur-
faces application, organic synthesis, water splitting, disinfection, anti-cancer therapy and water
purification (Fujishima et al., 2000). Among these, the later process has been most extensively
researched. The process is referred as heterogeneous, as the reaction medium consists of two
phases namely the solid (catalyst) and liquid (reagent). An ideal photocatalyst should be chem-
ically and biologically inert, photoactive, photostable, inexpensive, non-toxic, and should be
excited with solar light. Despite the existence of various chalcogenide semiconductor photo-
catalysts (oxides and sulfides), most studies involve TiO2 as semiconductor candidate despite
Decontamination of water by solar irradiation 5

its limitations, such as a low efficiency and a narrow light response factor due to its band-
gap (around 3.2 eV, corresponding to light <390 nm). The most prevalent applications of TiO2
areas white pigments in paints, plastics, paper, fibers, foods, pharmaceuticals, and personal care
products, mainly due to its light scattering properties and high refractive index (Skojaj et al.,
2011). The most prominent TiO2 photocatalyst, Degussa P25 (now known as AEROXIDE®TiO2
P25) has demonstrated good performance in photocatalytic applications. Although TiO2 P25 has
been widely used as a benchmark photocatalyst, its effectiveness has been limited by poor light
absorption in the visible region, as a result of its large band gap.
Mechanistic processes of TiO2 induced photocatalytic degradation of organic pollutants are
well described in the literature (Herrmann, 2010). Photocatalysis occurs due to absorption of a
photon with sufficient energy either equal or higher than the band gap energy (difference between
the valence band and the conduction band of the semiconductor) of the catalyst. When TiO2
is excited with light, an electron is promoted from the valence band to the conduction band,
generating a hole (h+ ) in the valence band and an electron (e− ) in the conduction band. The
photogenerated charge carriers can either migrate to the surface of the TiO2 , where it can actively
perform its role in the oxidation-reduction reactions with the pollutant or recombine in the bulk
or on the TiO2 surface. A generated photohole can react with an adsorbed water molecule or OH−
anions to form powerful HO• radicals (Eq. 1.1 and 1.2). When molecular oxygen is available, it
is adsorbed onto the surface of TiO2 and can scavenge an electron to form the superoxide radical
anion (O•−2 , Eq. 1.3).
h+ + H2 O → TiO2 + HO• + H+ (1.1)
+ − •
h + HO → TiO2 + HO (1.2)

e + O2 → TiO2 + O•−
2 (1.3)

However, a limitation of solar photocatalysis is the poor overlap of the solar spectrum with
the absorption spectrum of TiO2 (<5%). TiO2 doping with non-metals, metal-ion implantation,
codoping and sensitizingTiO2 with dyes have been applied to address this limitation and to improve
TiO2 photocatalytic efficiency. Doping results in a better absorption in the visible region as it
induces a change in the optical response of TiO2 and produces higher photonic yields. In contrast,
doping significantly increases the cost of the photocatalyst due to expensive ion implantation
facilities (Zhanget al., 2009).

1.2.2 Solar photo-Fenton


Fenton and Fenton-like processes are probably among the advanced oxidation processes most
applied for the treatment of industrial wastewater (Suty et al., 2004). The first proposals for
wastewater treatment applications were reported in the 1960s. Yet it was not until the early 1990s
that the first studies on the application of the photo-Fenton process for the treatment of wastewater
were published (Pignatello et al., 2006). Much of the literature on photo-Fenton includes the
possibility of driving the process with solar radiation because it seems to be the most suitable of
all AOPs to use sunlight, because soluble iron-hydroxy and especially iron-organic acid complexes
absorb part of the visible light spectrum, and not only UV radiation (Malato et al., 2009).
Hydrogen peroxide is decomposed to water and oxygen in the presence of iron ions in the
Fenton reaction in aqueous solutions (Eq. 1.4), as first reported by Fenton (1894). Mixtures of
ferrous iron and hydrogen peroxide are called Fenton reagents. Equations (1.4)–(1.6) show the
basic reactions in the absence of other interfering ions and organic substances. Regeneration of
ferrous iron from ferric iron by Equations (1.5) and (1.6) is the rate-limiting step in the catalytic
iron cycle, if iron is added in small amounts.
Fe2+ + H2 O2 → Fe3+ + OH− + HO• (1.4)
• +
Fe3+
+ HO 2 → Fe 2+
+ O2 + H (1.5)
Fe 3+
+ O•−
2 → Fe 2+
+ O2 (1.6)
6 S. Malato et al.

If organic substances (quenchers, scavengers, or in the case of wastewater treatment, pollutants)


are present in the Fe2+ /Fe3+ /H2 O2 system, they react in many ways with the generated hydroxyl
radicals. The generated organic radicals continue reacting, prolonging the chain reaction and
thereby contributing to reduce the consumption of oxidants in wastewater treatment by Fenton
and photo-Fenton. In aromatic pollutants, the ring system is usually hydroxylated before it is
broken up during oxidation, typically into intermediate degradation products containing quinone
and hydroquinone structures. In any case, sooner or later, ring opening reactions further mineralize
the molecule. One important drawback of the Fenton method, especially for total mineralization
of organic pollutants, is that carboxylic intermediates cannot be further degraded. Carboxylic and
dicarboxylic acids (L: mono- and dicarboxylic acids) are known to form stable iron complexes,
which inhibit the reaction with peroxide (Kavitha and Palanivelu, 2004). Hence, the catalytic iron
cycle reaches a arrest still before total mineralization is accomplished (Eq. 1.7).
H2 O2 ,dark
Fe3+ + nL → Fe Ln x+ −−−−−−→ no further reaction (1.7)

The primary step in the solar photoreduction of dissolved ferric iron is a ligand-to-metal charge-
transfer reaction in which intermediate complexes dissociate as shown in reaction (1.8). The ligand
may be any Lewis base able to form a complex with ferric iron (OH− , H2 O, HO− − −
2 , Cl , R-COO ,
R-OH, R-NH2 , etc.). Depending on the reacting ligand, the product may be a hydroxyl radical
such as in Equation (1.9) or other radical derived from the ligand. The direct oxidation of an
organic ligand is possible as well, as shown for carboxylic acids in Equation (1.10).

Fe3+ L + hυ → [Fe3+ L] → Fe2+ + L• (1.8)

Fe(OH)2+  + hυ → Fe2+ + HO• (1.9)

Fe(OOC − R)2+  + hυ → Fe2+ + CO2 + R• (1.10)


The ferric iron complex has different light absorption properties depending on the ligand; there-
fore, Equation (1.8) takes place with different quantum yields and also at different wavelengths.
Consequently, pH plays a crucial role in the efficiency of the photo-Fenton reaction, because it
strongly influences which complexes are formed. Thus, pH 2.8 has frequently been postulated as
optimum for photo-Fenton treatment, because there is no precipitation yet and the predominant
iron species in solution is [Fe(OH)]2+ , the most photoactive ferric iron-water complex (Eq. 1.9).
In fact, as shown in its general form in Equation (1.8), ferric iron can form complexes with many
substances and undergo photoreduction. Carboxylic acids are of special importance because they
are frequent oxidation intermediate products, and ferric iron-carboxylate complexes may have
much higher quantum yields than ferric iron-water complexes.
Fe3+ complexes present in mild acidic solutions absorb an appreciable amount of light in the
UV and in the visible region, and may complex with certain target compounds or their by-products.
These complexes typically have higher molar absorption coefficients in the near-UV and visible
regions than aquo complexes. Polychromatic quantum efficiencies from 0.05 to 0.95 are common
in the UV/visible range (Pignatello et al., 2006), making the photo-Fenton process suitable for
being driven by sunlight.

1.3 SOLAR TECHNICAL ISSUES

1.3.1 Hardware for solar AOPs


The specific hardware needed for solar photocatalytic applications have much in common with
those used for thermal applications. As a result, both reactors and photocatalytic systems have
followed at the beginning of their development conventional solar thermal collector designs,
such as parabolic troughs and non-concentrating collectors (Dillert et al., 1999). At this point,
their designs begin to diverge (Malato et al., 2002), since: (i) the fluid must be exposed to UV
Decontamination of water by solar irradiation 7

Figure 1.3. Normal solar irradiance (I) on the surface of Earth (ASTM E891-87, air mass 1.5), main light
absorbing gases and light absorption of Fe3+ species and TiO2 .

solar radiation, and, therefore, the absorber must be UV-transparent, and ii) temperature higher
than ambient does not play any significant role in the photocatalytic process and, therefore, no
insulation is required. Contrary to solar thermal processes, which are based on the collection
of large quantities of photons of all wavelengths to reach a specific temperature range, solar
photocatalysis is based on the collection of only high-energy short-wavelength photons to promote
photoreactions (Fig. 1.3). Most solar photocatalytic processes use UV or near-UV sunlight and
up to 580 nm. Sunlight at wavelengths over 600 nm is normally not harvested nowadays in any
photocatalytic process.
The original solar photoreactor designs for photochemical applications were based on line-
focus parabolic-trough concentrators (PTCs). In part, this was a logical extension of the historical
emphasis on trough units for solar thermal applications. Furthermore, PTC technology was rela-
tively mature and existing hardware could be easily modified for photochemical processes. The
first outdoors engineering-scale reactors developed were a converted solar thermal parabolic-
trough collector in which the absorber/glazing-tube combination has been replaced by a simple
Pyrex glass tube through which contaminated water could flow (Goswami, 1997). The main
disadvantages of PTCs were described in late 1990s (Alfano et al., 2000): (i) they can use
only direct radiation, (ii) they are expensive and (iii) they have low optical (solar tracking/light
reflection/concentration) and quantum efficiencies. The main conclusion was that solar con-
centrating devices (as PTCs) should be disregarded for photocatalysis. In middle 1990s, the first
non-concentrating collectors for photocatalysis were described as, in principle, they were cheaper
than PTCs as they have no moving parts or solar tracking devices. An extensive effort in the design
of small non-tracking collectors resulted in the design and testing of different non-concentrating
solar reactors (Bahnemann, 2004). However, the design of a robust non-concentrating photore-
actor is not trivial, since it must be weather-resistant, chemically inert and UV-transmissive. In
addition, flat plate non-concentrating systems require significantly more photoreactor area than
concentrating photoreactors and, as a consequence, full-scale systems (normally composed of
hundreds of square meters of collectors) must be designed to withstand the operating pressures
anticipated for fluid circulation through a large field. In uncovered, non-concentrating systems
exposed to the ambient, reactants and catalyst could become contaminated (Fig. 1.4). Very often,
the chemical inertness of the materials used (to resist corrosion caused by outdoor operation and
exposure to solar irradiation) for constructing the flat plate non-concentrating collector would be
difficult to guarantee. Consequently, the use of these photoreactors was abandoned by the main
research groups.
8 S. Malato et al.

Figure 1.4. Example of non-concentrating collector developed during the 1990s: water-fall for photo-
Fenton applications.

To design a solar collector for photocatalytic purposes, there is a group of constraints that
should be fulfilled: (i) the effective collection of UV-vis radiation, (ii) working temperatures as
close as possible to ambient temperature in order to avoid loss of volatile organic compounds, (iii)
to take into account that quantum efficiency decreases with irradiance, (iv) to take into account
that concentrators collect 1/RC (concentration factor, RC, ratio between radiation collection area
and photoreator area) of the available diffuse radiation. As a result of these considerations, the
concentration for photocatalytic applications should be RC = 1. Finally, its construction should be
economical and easy, and the reactor should maintain a low pressure drop. As a consequence, the
use of tubular photoreactors has a decisive advantage because of the inherent structural efficiency
of tubing. Tubing is also available in a large variety of materials and sizes and it is a natural choice
for a pressurized fluid system.
There is a category of low concentration collectors, called Compound Parabolic Concentrators
(CPCs), which are used in thermal applications. They are an interesting option between parabolic
concentrators and static flat systems, since they combine characteristics of both: they concentrate
solar radiation, but they conserve the properties of the flat plate collectors, being static and
collecting diffuse radiation. Thus, they also constitute a good option for solar photochemical
applications (Ajona and Vidal, 2000). CPCs are static collectors with surface following an involute
around a cylindrical reactor tube and designed with a RC = 1; they have advantages of both PTCs
and flat plate collectors. The light reflected by the CPC is distributed all around the tubular
receiver, so that almost the entire circumference of the receiver tube is irradiated and the light
incident on the photoreactor is the same that would impinge on a flat plate. CPCs can make
highly efficient use of both direct and diffuse solar radiation, without the need for solar tracking.
There is no evaporation of possible volatile compounds because water does not heat up. They
have high optical efficiency, since they make use of almost all the available radiation, and high
quantum efficiency, as they do not receive a concentrated flow of photons. Flow also can be easily
maintained turbulent inside the tube reactor. They are considered nowadays the state of the art
for solar collectors in photocatalysis (Colina-Márquez et al., 2010). As in a parabolic trough,
the water is more easily piped and distributed than in many one-sun designs. CPC reflectors
are usually made of polished aluminum and the structure can be a simple photoreactor support
Decontamination of water by solar irradiation 9

Figure 1.5. Close view of CPC shape with tubing.

frame with connecting tubing (Fig. 1.5). Since this type of reflector is considerably less expensive
than tubing, their use is very cost-effective compared to deploying non-concentrating tubular
photoreactors without use of any reflectors.
In addition to the solar collector type, the most important technical issues related with solar
photocatalysis hardware are the reflective surface and the reactor tube (Blanco and Malato, 2003).
For solar photocatalytic applications, the reflective surface must clearly be made of a highly reflec-
tive material for UV radiation. The reflectivity (reflected radiation/incident radiation) between
300 and 400 nm of traditional silver-coated mirrors is very low and aluminum-coated mirrors is
the best option in this case. Aluminum is the only metal surface that is highly reflective throughout
the UV spectrum. The photocatalytic reactor must contain the catalyst and be transparent to UV
radiation providing good mass transfer of the contaminant to an irradiated photocatalyst surface
with minimal pressure drop across the system. Adequate flow distribution inside the reactor must
be assured. The choice of materials that are both transmissive to solar UV light and resistant to its
destructive effects is limited. Common materials that meet these requirements are fluoropolymers
and several types of glass. Quartz has excellent UV transmission as well as good temperature
and chemical resistance, but high cost. Fluoropolymers are a good choice due to their good UV
transmittance, excellent UV stability and chemical inertness. Glass is another alternative for pho-
toreactors. Borosilicate glass has good transmissive properties in the solar range with a cut-off of
about 285 nm. Therefore, such a low-iron-content glass would seem to be the most adequate. If a
large field is being designed, large collector area means also a considerable number of reactors
and, as consequence, high pressure drop. Thus, fluoropolymer tubes are not the best choice of
material since high-pressure is linearly related to thickness and could result in higher cost and
lower UV transmittance.
One of the most important parameters in a tubular photoreactor design is the diameter, as in
both homogeneous or heterogeneous photocatalysis it must be guaranteed that all arriving useful
photons are kept inside the reactor and do not go through it without intercepting the catalyst. The
intensity of irradiation affects the relationship between reaction rate and catalyst concentration.
The dispersion and absorption of light causes photon density to diminish over the length of the
optical path within a catalyst suspension. It is also important to take into account that at higher
irradiance, the catalyst concentration could be higher. The absorption of radiation in a photocat-
alytic tubular reactor is influenced by the geometry of the reactor and by the optical properties of
the photocatalyst. The modeling of the radiation absorbed in solar photoreactors (tubular reactors
and compound parabolic collectors) has been reported by several authors (Cassano and Alfano,
2000; Colina et al., 2010; Giménez et al., 1999; Malato et al., 2004).
Practical inner diameters for tubular photoreactor would be in the range of 25 to 50 mm. Very
small diameters are not convenient because of the associated high pressure-drop, and very large
diameters imply a considerable dark volume, thus reducing the overall system efficiency. Use of
10 S. Malato et al.

wavelengths where the catalyst does not absorb is better for determining the optimum catalyst
concentration as a function of light-path length. Under these conditions, measurement of photon
losses is only affected by turbidity and it is easier to evaluate the effect of the photoreactor diameter
(Fernández et al., 1999). After many experiments and simulations with different photoreactors,
the optimum TiO2 concentration obtained with sunlight is around a few hundreds of mg L−1 with
a photoreactor diameter in the range of 25 to 50 mm, as said before. This statement corresponds
to heterogeneous photocatalysis, but it is interesting to note than nowadays applications related
with photo-Fenton are also very usual. In this case, it is important to take into account that
Fe3+ (related species and organic complexes, see former section) absorbs solar photons, and
this effect must be considered when determining the optimum load as a function of light-path
length in the photoreactor. Iron optimal concentration is affected more by light-path length than
the TiO2 process. As these calculations are strongly affected by different complexes formed
by iron (III) during the photo-Fenton process, they had been experimentally demonstrated. The
optimum concentrations of 0.2–0.5 mmol L−1 were obtained after many experiments with different
photoreactors and different wastewaters under sunlight (Cassanoet al., 2011; Gernjak et al., 2006;
Lucas et al., 2009; Oller et al., 2007; Sirtori et al., 2009; Zapata et al., 2010).

1.3.2 Solar photocatalytic treatment plants


The first outdoor engineering-scale reactor developed was a converted solar thermal parabolic-
trough collector in which the absorber/glazing-tube combination had been replaced by a simple
Pyrex glass tube through which contaminated water could flow (Goswami, 1997). Since that time,
research all over the world has proposed a number of reactor concepts and designs, including
concentrating and non-concentrating reactors (Malato et al., 2009). The design procedure for a
solar photocatalytic system requires the selection of a reactor, catalyst operating mode (slurry
or fixed matrix), reactor-field configuration (series or parallel), treatment-system mode (once-
through or batch), flow rate, pressure drop, pretreatment, catalyst and oxidant loading method,
pH control, etc. Usually, a photocatalytic plant is constructed with several solar collectors. All the
modules are connected in series or parallel, but with valves that permit to bypass any number of
them. Centrifugal pumps with an electric motor (calculated to provide sufficient flow when the
maximum length of the system is used) have to be installed to move the treatment water through
the reactor. Either a flow-rate control loop made up of a flow meter connected to a controller,
which in turn governs an automatic electric valve, or an electric pump with a speed controller has
to be installed to regulate the flow to the rate desired. The most important sensors required for
the system are temperature, pressure and dissolved oxygen (at least in the reactor outlet). Other
sensors, such as pH, selective electrodes, etc., could be useful depending on the type of wastewater
to be treated. A UV-radiation sensor must be placed in a position where the solar UV light reaching
the photoreactor can be measured, permitting the evaluation of the incident radiation as a function
of hour of the day, clouds, atmospheric or other environmental variations. Solar photocatalytic
plants are frequently operated in a recirculating batch mode. The fluid is continuously pumped
between a reactor zone and a tank in which no reaction occurs, until the desired degradation is
achieved. The systems are operated in a discontinuous manner by recirculating the wastewater
with an intermediate reservoir tank and centrifugal pump.
Under the “SOLARDETOX” project (Solar Detoxification Technology for the Treatment of
Industrial Non-Biodegradable Persistent Chlorinated Water Contaminants), a Consortium (coor-
dinated by Plataforma Solar de Almería, Spain) was formed in the EU for the development and
marketing of solar photocatalytic plants. The main goal (financed by the EC-DGXII through
the BriteEuram III Program, 1997–2000) was to develop a commercial non-concentrating solar
detoxification system using CPC. Based on accumulated experience in pilot plant design, con-
struction and testing (Blanco et al., 2000), a full-size demonstration plant (Fig. 1.6.) was erected
to treat 1 m3 of water contaminated. The main plant characteristics were: (i) 2 rows of 21 collectors
each; (ii) total collector (tilted 40◦ ) aperture area = 98 m2 ; (iii) total loop volume = 675 mL; (iv)
total plant volume = 975 L; (v) 200 mg L−1 TiO2 slurry; (vi) 1.5 × 1.5 m collectors with sixteen
Decontamination of water by solar irradiation 11

Figure 1.6. Full-size demonstration plant for treating wastewater with slurries of TiO2 .

29.2-mm-I.D. tubes. A complete module is formed by a series of collectors connected in a row. The
final prototype plant consists of E-W oriented parallel rows of 21 collectors each. The structure
was slightly tilted (1%) in the same direction to dry out rain water and avoid its accumulation in
the CPC troughs. As this plant was to be a demonstration of what a commercial plant would be
like, operation was fully automatic. Among the electronic control devices, the most useful was
a solar UV-A sensor that integrated solar UV with time during the treatment. This sensor was
connected to a controller and, once sufficient energy for finishing the treatment has been achieved,
it stopped the main pump, indicating to the operator that the treatment has been completed. The
plant was operated in batch mode at flow rates between 30 to 35 m3 h−1 , assuring fully turbulent
flow and good mixing. Other photocatalytic plants have been installed since 2000 and reviewed
(Bahnemann, 2004; Chong et al., 2010; Malato et al., 2009).

1.4 TREATMENT OF INDUSTRIAL WASTEWATERS

Biological degradation of a chemical refers to the elimination of the pollutant by the metabolic
activity of living organisms, usually microorganisms and, in particular, bacteria and fungi that
live in natural water and soil. In this context, conventional biological processes do not always pro-
vide satisfactory results, especially for industrial wastewater treatment, since many of the organic
substances produced by the chemical industry are toxic or resistant to biological treatment. There-
fore, the only feasible option for such biologically persistent wastewater is the use of advanced
technologies based on chemical oxidation, such as the advanced oxidation processes (AOPs),
widely recognized as highly efficient treatments for recalcitrant wastewater. Chemical oxidation
for complete mineralization is generally expensive because the oxidation intermediates formed
during treatment tend to be more and more resistant to their complete chemical degradation. One
attractive potential alternative is to apply these chemical oxidation processes in a pretreatment to
convert the initially persistent organic compounds into more biodegradable intermediates, which
would then be treated in a biological oxidation process with a considerably lower cost. Studies
have long shown that the biodegradability of a waste stream changes when subjected to prior
chemical oxidation (Pulgarín et al., 1999; Rizzo, 2011). Therefore, the main role of the chemi-
cal pretreatment is partial oxidation of the biologically persistent part to produce biodegradable
reaction intermediates. The percentage of mineralization should be minimal during the pretreat-
ment stage in order to avoid unnecessary expenditure of chemicals and energy, thereby lowering
the operating cost. However, if the pretreatment time is too short, the reaction intermediates
generated could still be structurally very similar to the original non-biodegradable and/or toxic
components. Experimental examples of sequential chemical and biological oxidation treatment
have been previously reviewed by Scott and Ollis (1995; 1997), Mantzavinos and Psillakis (2004),
Augugliaro et al. (2007), and others.
12 S. Malato et al.

Appropriate techniques must be combined to provide technically and economically feasible


options. The performance of an AOP treatment could be enhanced in several ways. The first
possibility is to position the AOP in a sequence of physical, chemical and biological treatments.
Such a treatment approach often involves at least one AOP step and one biological treatment
step. Whether the AOP or the biological process is first in the treatment line, the overall purpose
of reducing costs will be nearly the same as minimizing AOP treatment and maximizing the
biological stage, because of the wide difference in the cost of the two treatments. The key issue is
to design the process for the best overall economic and ecological performance. Measurement of
the combined process efficiency depends on the purpose of the treatment, but it normally requires
the independent optimization of each chemical and biological step. For example, the extent of
mineralization of the organic compounds may be a measure of efficiency if highly pure water or
an effluent with a specific dissolved organic carbon limit is needed. The main purpose of other
treatments may be reduction of toxicity or elimination of a specific pollutant. Determining the
target is an essential step in combination studies, since it helps to define process efficiency and
provides a basis to compare the different operating conditions and to optimize the process.
Calculation of the individual biological and chemical oxidation efficiencies is important to
find the optimal operating conditions for the combined process. This task involves a profound
knowledge of both biological and chemical processes. Therefore, several analytical parameters
must be monitored during each step of the treatment line. Chemical parameters normally mea-
sured are the total organic carbon (and/or chemical oxygen demand), the concentration of specific
pollutants which could be present in the target wastewater (by chromatographic methods, such as
HPLC-UV), and complete oxidation of heteroatoms released (Cl, N, P, . . .) as inorganic species
(Cl− , NO− 3−
3 , PO4 , . . .) into the media from contaminants completely degraded during the AOP
treatment (by ion chromatography or commercial tests designed for each specific ion). Regarding
biological assays, it is very important to perform toxicity analyses (with organisms like Vibrio
fischeri, Daphnia magna, activated sludge by respirometric assays, etc.) and biodegradability
tests (using activated sludge) to ensure the conditions of the AOP effluent to be subsequently
treated by a conventional biodegradation process. In the biological system itself, and apart from
daily control analyses such us total suspended solids and volatile solids, total organic carbon
and chemical oxygen demand, pH and dissolved oxygen in the system, etc., the measurement of
anions and cations present in the biological media is also essential since, on one hand, nutrients
are vital to the micoorganism populations in the activated sludge, and, on the other hand, moni-
toring the nitrogen species provides much information related to nitrification and denitrification.
This series of analytical parameters satisfies the engineering needs for designing the coupling
strategy.

1.4.1 Toxicity and biodegradability assessment


Toxicity analysis of the wastewater during different stages of its AOP treatment is done by acute
toxicity testing (with 96 hours of maximum exposure time) using different microorganisms. There
are many available procedures for toxicity bioassays; nevertheless, as toxicity is a biological
response, there is no universal monitoring system and, therefore, to increase the reliability of
toxicity assays, different organisms representative of taxonomic groups from typical of the local
environment must be employed (Rizzo, 2011).
In some cases, it has been found that the toxicity of the original effluent grows during the early
pretreatment up to a maximum due to the formation of toxic intermediates. It is important to keep
in mind that such reaction intermediates formed during chemical oxidation could be more toxic to
the biological systems than the original compound, and that different oxidation processes can lead
to different intermediates. On the other hand, the oxygen demand obtained in respirometric assays
have recently turned into an excellent control parameter as it represents a direct measure of the
correct activity and viability of microorganisms present in aerobic activated sludge. Furthermore,
as this test represents a direct assessment of the primary function in a process based on activated
sludge, it can be used as an efficient tool for the measurement of acute toxicity that could provoke
Decontamination of water by solar irradiation 13

Figure 1.7. Solar detoxification of wastewater according to three different methods vs. illumination time.

different industrial wastewater inlet on the activated sludge of a municipal wastewater treatment
plant (MWWTP) (Vilar et al., 2009).
In general, toxicity assays do not require a strong investment in equipment or excessively
specialized training in their handling. But, it is also very important to mention that previously to
the application of these toxicity analyses, any toxic substance that could be present during the
chemical pretreatment (hydrogen peroxide, high amounts of catalysts, etc.) must be eliminated
from the media and the pH of the water must be kept between 6.5 and 7.5. Most AOPs lower the pH
due to the generation of inorganic acids, or they need to operate at a certain pH (e.g., pH around
3 for Fenton), and this is why prior neutralization is required for toxicity and biodegradability
tests, and for a final biological treatment step.
An example is shown in Figure 1.7, where toxicity of samples taken after different periods
of the photo-Fenton process applied to wastewater was estimated according to the inhibition of
the oxygen uptake rate (OUR) of activated sludge. Although the inhibition of the OUR could
be considered as a good parameter to estimate the acute toxicity of chemicals towards activated
sludge, it is not able to detect the mid and long time effect of toxicants due to the short time that
the micro-organisms are exposed to the pollutant. Thus, a complementary method, more time
consuming, based on the measurement of cumulative oxygen consumption, appears necessary to
ensure detoxification. For this purpose, the inhibition of the activity of the sludge was determined
by measuring the BOD5 of a very biodegradable mixture of glucose and glutamate in the presence
of studied effluent. Figure 1.7 shows that the initial toxicity obtained according to this method
was higher than that measured in the short time experiments. However, similar qualitative trends
can be observed. At the end of treatment, negative inhibitions (stimulation) were determined.
Inhibition of the luminescence of V. fischeri was also evaluated. The toxicity obtained according
to this method was much higher, and, although a decrease was observed along the treatment, a
significant value remained. The higher sensibility of this method can be attributed to the specific
nature of the organism (V. fischeri) involved in the measurement, as the activated sludge consists
in a consortium of microorganisms present in the biological reactors of sewage treatment plants,
and that have suffered adaptation to the effect of a wide range of toxicants. However, combining
14 S. Malato et al.

data obtained from all the three series of bioassays and chemical analyses, it can be stated that,
although complete remediation of the effluent was not achieved, it might be compatible with a
biological treatment after elimination of the pesticides.
When considering combined chemical oxidation and biological processes for treating recal-
citrant contaminants, biodegradability assessment is required not only of the raw wastewater,
but also during the AOP pretreatment (Ballesteros et al., 2010). In this sense, enhancement of
biodegradability by an AOP application can be monitored by means of (Sarria et al., 2003):
• Analysis of general parameters, such as the biological oxygen demand (BODx ), chemical
oxygen demand (COD), and dissolved organic carbon (DOC).
• Evolution of the BOD5 /COD ratio or the average oxidation state (AOS). These ratios provide
an approximate index of the proportion of organic substances present in the wastewater that
are biodegradable under aerobic conditions (Sarria et al., 2002).
• Long activated sludge biodegradability assays, such as the Zahn-Wellens test, which is used
to evaluate the biodegradability of water-soluble, non-volatile organic contaminants when
exposed to a relatively high concentration of microorganisms. It takes around 28 days kept at
20–25◦ C under diffuse illumination (United States Environmental Protection Agency, 1996).
• Oxygen uptake rate by respirometric measurements (short analysis). Respirometry measures
the oxygen used by bacteria during growth, which is interrelated to reduction in BOD. In
this biological assay, the oxygen uptake rate from a mixture of raw or pretreated wastewater
and activated sludge (in endogenous phase with autotrophic bacteria activity inhibited) is
measured for a contact period of around 20 minutes. The different wastewater COD fractions
(biodegradable, non-biodegradable, non-soluble, etc.), can also be determined by respirometric
assays (Lagarde et al., 2005).

1.4.2 Industrial wastewater treatment by combined AOPs/biotreatment


Industrial wastewater characteristics vary not only with the industry that generates them, but also
within the industry. These characteristics are also much more diverse than domestic wastewater,
which is usually qualitatively and quantitatively similar in its composition. On the contrary,
treatment of industrial wastewater is a complex problem due to the wide variety of compounds
and concentrations it may contain. In this sense, Scott and Ollis (1995) have identified four types
of wastewater as potentially treatable by combined AOPs/biological degradation: wastewater
containing biorecalcitrant compounds such as large macromolecules like soluble polymers that
are not easily biodegradable due to their large size and lack of active centers; highly biodegradable
industrial wastewater which still requires chemical post treatment as it contains a large amount of
biodegradable organic compounds in addition to small concentrations of recalcitrant compounds;
wastewater containing inhibiting compounds which are somewhat toxic to a certain percentage of
some biological cultures; wastewater containing inert intermediates such as specific metabolites
which must be effectively degraded or, they would accumulate in the medium and inhibit growth
of the microorganisms.
Among wastewater containing inhibiting compounds to biological cultures are soluble pes-
ticides, which are a serious threat to surface- and groundwater, since their high solubility in
water makes extremely easy their propagation in the environment. Most pesticides are resistant
to chemical and/or photochemical degradation under typical environmental conditions. Part of
the larger-scale contamination is known to result from non-agricultural uses of pesticides (Skark
et al., 2004) or from point sources, including discharge from farmyards following filling and
washing activities (Neumann et al., 2002). Nevertheless, diffuse contributions of pesticides to
water are also important. These come predominantly from applications of pesticides including
spray drift, surface runoff and leaching to field drains. Less significant routes to surface water
include groundwater seepage, subsurface lateral flow and wet or dry deposition following longer
range transport in air. Movement via field drains has been shown to be important in a number
of countries and rapid transport of pesticide residues in drain flow has been demonstrated in a
Decontamination of water by solar irradiation 15

large number of field experiments (Brown and van Beinum, 2009). Pesticide contamination in
wastewater from these sources may be as high as 500 mg L−1 . However, as biological methods are
normally susceptible to such toxic compounds, which inactivate waste-degrading microorganisms,
a potentially useful approach is to partially pretreat the toxic waste by oxidation technologies,
producing intermediates that are more readily biodegradable. Many oxidation treatments have
traditionally been studied for this purpose, solar photocatalysis being highly successful. In any
case, the efficiency of these methodologies has hardly been assessed under real conditions, i.e.,
in the presence of a mixture of several pesticides and their formulating agents at concentrations
over 100 mg L−1 (Zapata et al., 2010).
Pharmaceutical residues are another group of compounds of particular interest and unknown
fate. For instance, during monitoring in Italy, France, Greece and Sweden, carbamazepine,
clofibrate, phenazone and aminopyrine, clofibric acid, diclofenac, fenofibrate, fenoprofen, flur-
biprofen, gemfibrozil, ibuprofen, ketoprofen and naproxen, all belonging to different therapeutic
pharmaceuticals classes, were found in the effluents of sewage treatment plants (Andreozzi et al.,
2003). In general, the presence of residual pharmaceuticals in the environment and in aquatic
systems in particular constitutes a serious problem as they are extremely resistant to biologi-
cal degradation and usually escape intact from conventional treatment plants. They may have
serious toxic and other effects on humans and other living organisms, and they are present at
minute concentrations, thus requiring more sophisticated and laborious analytical tools for their
accurate determination. Solar photo-Fenton has been used in combination with an aerobic bio-
logical system for the treatment of pharmaceutical wastewater. For example, a combined solar
photocatalytic-biological pilot plant system was employed to enhance the biodegradability and
complete mineralization of a biorecalcitrant industrial compound (α-methylphenylglycine, a com-
mon pharmaceuticals precursor), dissolved in distilled water and simulated seawater at 500 mg L−1
(Oller et al., 2007). Evaluation of the combined AOP/biological system developed demonstrated
that in batch mode operation, photo-Fenton pretreatment completely removed the pollutant and
enhanced its biodegradability, producing a biocompatible effluent, which was completely mineral-
ized by the biological system in an immobilized biomass reactor. Solar photo-Fenton has also been
employed for the treatment of a real pharmaceutical wastewater (TOC = 775 mg L−1 ) containing a
non-biodegradable antibiotic pertaining to the quinolone group called nalidixic acid (45 mg L−1 ),
It was demonstrated that the more suitable combination was biotreatment/photo-Fenton, as the
industrial wastewater was highly biodegradable but it still required chemical posttreatment for
removing biorecalcitrantnalidixic acid (Sirtori et al., 2009).
The textile industry is very water intensive. Water is used for cleaning the raw material and
for many flushing steps throughout production. Textile wastewater includes additions of a wide
variety of dyes and chemicals that make the chemical composition of textile industry efflu-
ents an environmental challenge. Major pollutants specifically found in textile wastewater are
suspended solids, highly stable chemical oxygen demand, dyes giving intense color and other
soluble substances (World Bank, 2007). Removal of color from textile industry and dyestuff man-
ufacturing industry wastewaters represents a major environmental concern. Its strongest impact
on the environment is related to primary water consumption (80–100 m3 /ton of finished textile)
and wastewater discharge (115–175 kg of COD/ton of finished textile, a large range of organic
chemicals, low biodegradability, color, and salinity) (Savin et al., 2008). Typical textile industry
wastewater characteristics can be summarized by a COD range from 150 to 12,000 mg L−1 , total
suspended solids between 2900 and 3100 mg L−1 , total Kjeldahl nitrogen from 70 to 80 mg L−1 ,
and BOD range from 80 to 6000 mg L−1 leading to a BOD/COD ratio of around 0.25, showing that
it contains large amounts of non-biodegradable organic matter. AOPs have been widely shown to
have the greatest promise for treating textile wastewater (Hai et al., 2007), the use of solar photo-
catalysis being relevant (Bousselmi et al., 2004; Singh and Arora, 2011). There are other industrial
wastewaters where solar AOPs have been applied, including paper mill wastewater (Amat et al.,
2005; Parilti and Akten, 2011; Sattler et al., 2004), olive mill wastewater (Baransi et al., 2012;
Gernjak et al., 2005; 2007), landfill leachate (Cassano et al., 2011; Wiszniowski et al., 2006),
winery and distillery wastewater (Lucas et al., 2009; Mosteo et al., 2007; Rodríguez et al., 2008).
16 S. Malato et al.

Table 1.1. Average of microcontaminants found in different secondary effluents from MWTP at
>250 ng L−1a .

Contaminant ng L−1 Contaminant ng L−1 Contaminant ng L−1

Caffeine 18530 Hydrochlorothiazide 2050 Antipyrine 830


4-AAA 13730 4-AA 1490 Isoproturon 710
Paraxanthine 6820 Naproxen 1390 Ciprofloxacin 700
4-FAA 6740 Diclofenac 1330 Acetaminophen 610
Nicotine 6530 Ofloxacin 1080 Diuron 540
Cotinine 6040 Atenolol 920 Ketoprofen 450
Ibuprofen 5300 Ranitidine 920 Trimethoprim 330
Gemfibrozil 3650 Codeine 890 Venlafaxime 330
Furosemide 2210 Sulfamethoxazole 840 Azithromycin 260
4-MAA 2090
a Analyses performed in different MWTP effluents revealed the presence of 62 microcontaminants at

concentrations ranging from 3 ng L−1 , but only those >250 ng L−1 are shown.

1.5 TREATMENT OF SECONDARY EFFLUENTS

As commented in the introduction, a large number of compounds in the group of unregulated


organic substances usually known as emerging contaminants (ECs) have been reported to be
possible environmental contaminants. Because they are ingredients in daily household products
and are only partially degraded during conventional wastewater treatment, municipal wastewater
treatment plant (MWTP) effluents have been pointed out as the main source of these pollutants in
the aquatic environment. Although the concentrations of these contaminants are in the µg-ng L−1
range, they have become a matter of concern due to their sheer numbers and continuous disposal
into the environment, and today represent a serious threat to the quality of receiving waters (rivers,
aquifers and groundwater) (Kümmerer, 2009; Pal et al., 2010). Treated wastewater from secondary
effluents is being widely reused, but the knowledge of the potential effects that such reuse might
induce, especially with regard to organic compounds in the treated effluents, is still incomplete
(Calderón-Preciado et al., 2011; Fatta-Kassinos et al., 2011). It is therefore important to degrade
these contaminants in the wastewater prior to their release into the environment, and moreover, if
it will be used for irrigation. Table 1.1 presents analyses performed in different MWTP effluents
showing an average of different samples. The analytical method used for the target compounds
was developed in a 3200 QTRAP MS/MS system (Biosystems, Concord, ON, Canada).The
Instrumental Detection Limits (IDLs) were in the 0.1–100 pg range. Relative standard deviations
for the samples were below 20% in all cases (Bueno et al., 2007).
Among the advanced treatment technologies for the degradation of these micro- and emerging
contaminants in water, advanced oxidation processes (AOPs) area particularly attractive option.
One of particularly interest is photo-Fenton; however, one major drawback of this type of treat-
ments is the need for a low pH (optimal pH 2.8), as said in Section 1.2.2., as iron precipitates at
higher pH (Pignatello et al., 2006). In addition, at high pH, the treated water has to be neutral-
ized before reuse, raising the salt content, which would be negative for certain purposes such as
irrigation.
Results shown (Table 1.2, Fig. 1.8) compares two different photo-Fenton processes, conven-
tional photo-Fenton at pH 2.8 and modified photo-Fenton at neutral pH with minimal amount of
Fe (5 mg L−1 ) and minimal initial H2 O2 concentration (50 mg L−1 ). As Fe precipitates at neutral
pH, complexing agents that: (i) are able to form photoactive species (Fe3+ L, Eq. 1.8), (ii) do not
pollute the environment, and (iii) do not reduce the biodegradability of the water, have to be used
to keep iron in solution. As MWTP effluents do not normally contain these substances, since they
are removed during water potabilization (main source of water in MWTPs) or are biodegradable
Decontamination of water by solar irradiation 17

Table 1.2. Initial and residual concentration of the remaining contam-


inants in MWTP effluents treated with photo-Fenton at pH
3 and 5 mg L−1 Fe after 50 minutes of solar irradiationa .

Contaminant C0 , µg L−1 CF , µg L−1

Nicotine 43.0 1.2


Paraxanthine 9.5 0.16
Caffeine 44.0 0.10
4-AAA 37.0 0.10
4-FAA 8.0 0.02
Cotinine 15.0 3.7
Diuron 2.4 0.001
 of all contaminants 180.0 5.3
a Analyzesperformed in this MWTP effluent revealed the presence of
45 microcontaminants.

(and therefore consumed during the secondary treatment step), they have to be added during the
ternary treatment.
Humic substances (HS) and in particular humic acids (HA) are naturally occurring organic sub-
stances that result from microbiological and chemical transformation of organic debris. Although
HS differ depending on the source, some general properties are similar. They are the largest frac-
tion of dissolved organic matter (NOM) in natural water, have strong light absorbance properties,
and generate excited triplet states (3 HS*), various reactive oxygen species as singlet oxygen
(1 O2 ), and hydroxyl radicals. They behave like colloids, and have absorptive qualities. They con-
tain carboxylic acids, phenolic, alcoholic quinine, amino and amido groups that enable them to
support ion exchange and redox processes, form stable complexes, and stabilize free radicals
(Lipczynska-Kochany et al., 2008).
Table 1.2 shows the experiments conducted with photo-Fenton at pH 2.8 and 5 mg L−1 Fe.
The DOC changed from 26 to 17 mg L−1 , H2 O2 consumption was 80 mg L−1 at the end of the
experiment after illumination time of 50 minutes, and Fe concentration was stable at 5 mg L−1 .
The initial concentration of all contaminants (45 contaminants were detected in this specific
effluent from MWTP secondary treatment) was extremely high, with 180 µg L−1 , and this number
declined to 5.3 µg L−1 , deriving to a degradation percentage of 97%. The high load of certain
substances like caffeine, cotinine, paraxanthine and the two metabolites 4-AAA and 4-FAA
(dipyronemetabolites) could be explained through the higher seasonal use in autumn (beginning
of the cold season). All other contaminants degraded below their limit of detection of the analytical
method within the first minutes after irradiation.
When treating MWTP effluents with a modified photo-Fenton system at neutral pH with HA,
the results were a little more disappointing. One experiment conducted with 10 mg L−1 HA had
an initial pH of 6.5 and a final pH of pH 4.7. The experimental time was 144 minutes, DOC was
stable around 16 mg L−1 , 57 mg L−1 H2 O2 was consumed. Out of the initial 46 contaminants
detected with an initial concentration of 601 µg L−1 , 14 contaminants were still present with a
concentration of 9.8 µg L−1 (1.6% of the initial concentration) at the end of the experiment (see
Fig. 1.8).
The degradation of the 46 present contaminants can be seen in Figure 1.8 and are ordered by
concentration to make them more visible. Contaminant 39 is actually the sum of all contaminants
that are present in concentrations lower than 100 ng L−1 . Clarithromycin was still present at the end
of the experiment. When the treatment was performed with 5 mg L−1 HA under similar conditions,
after 135 minutes the concentration of contaminants was 31.0 µg L−1 , and 32 contaminants were
still present at the end of the experiment with concentrations between 0.1 and 7.6 µg L−1 , revealing
the importance of a proper selection of HA concentration.
18 S. Malato et al.

Figure 1.8. Solar detoxification of wastewater with a modified photo-Fenton system at neutral pH with
HA vs. illumination time. 1 Gemfibrozil, 2 caffeine, 3 ibuprofen, 4 cotinine, 5 4-AAA,
6 paraxanthine, 7 nicotine, 8 4-FAA, 9 atrazine, 10 naproxen, 11 diclofenac, 12 4-MAA,
13 hidrochlorothiazide, 14 furosemide, 15 sulfamethoxazole, 16 benzafibrate, 17 4-AA, 18
atenolol, 19 ofloxacin, 20 diuron, 21 codeine, 22 ranitidine, 23 ciprofloxacin, 24 ketoprofen,
25 trimethoprim, 26 antipyrine, 27 simazine, 28 erythromycin, 29 azithromycin, 30 velafaxime,
31 chlorfenvinphos, 32 sulfapyridine, 33 pravastatin, 34 fenofibric acid, 35 citalopram HBr,
36 lincomycin, 37 acetaminophen, 38 indomethacine, 39 (diazepam, propanolol, mepivacaine,
mefenamicacid, nadolol, norfloxacin, carbamazepine, clarithromycin).

1.6 CONCLUSIONS

AOPs are well known for their capacity for oxidizing and mineralizing almost any organic contam-
inant, and different approaches have been proposed based on the use of renewable energy sources,
i.e. sunlight as irradiation source. Solar TiO2 photocatalysis and photo-Fenton are low cost AOP
systems already developed at real scale (using solar compound parabolic concentrators) and also
suitable to be combined with biological processes. Investigation concerning such options would
undoubtedly contribute to enhance the applicability of combined AOP+Biological processes for
treating different industrial wastewaters.
Unregulated organic substances usually known as emerging contaminants have been reported
to be possible environmental contaminants. Municipal wastewater treatment plant effluents have
been pointed out as the main source of these pollutants in the aquatic environment. Among the
advanced treatment technologies for the degradation of these micro- and emerging contaminants
in water, one of particularly interest is solar photo-Fenton at neutral pH using complexing agents
with minimal amount of Fe and minimal H2 O2 .

ACKNOWLEDGEMENTS

The authors wish to thank the Spanish Ministry of Science and Innovation for its financial
assistance under the “Fotoreg” Project (Ref CTQ2010-20740-C03-02).

REFERENCES

Ajona, J. A. & Vidal, A.: The use of CPC collectors for detoxification of contaminated water: Design,
construction and preliminary results. Sol. Energy 68 (2000), pp. 109–120.
Decontamination of water by solar irradiation 19

Alfano, O.M., Bahnemann, D., Cassano, A.E., Dillert, D. & Goslich, R.: Photocatalysis in water environments
using artificial and solar light. Cat.Today 58 (2000), pp. 199–230.
Amat, A.M., Arqués, A., López, F. & Miranda, M.A.: Solar photo-catalysis to remove paper mill wastewater
pollutants. Sol. Energy 79 (2005), pp. 393–401.
Andreozzi, R., Raffele, M. & Nicklas, P.: Pharmaceuticals in STP effluents and solar photodegradation in
aquatic environment. Chemosphere 50 (2003), pp. 1319–1330.
Augugliaro, V., Litter, M., Palmisano, L. & Soria, J.: The combination of heterogeneous photocatalysis with
chemical and physical operations: A tool for improving the photo process performance. J. Photochem.
Photobio. C: Photochem. Rev. 7 (2007), pp. 123–144.
Bahnemann, D.: Photocatalytic water treatment: Solar energy applications. Sol. Energy 77 (2004),
pp. 445–459.
Ballesteros, M.M., Casas, J.L., Oller, I., Malato, S. & Sánchez, J.A.: A comparative study of different
tests for biodegradability enhancement determination during AOP treatment of recalcitrant toxic aqueous
solutions. Ecotoxicol. Environ. Safety 73 (2010), pp. 1189–1195.
Baransi, K., Dubowski, Y. & Sabbah, I.: Synergetic effect between photocatalytic degradation and adsorption
processes on the removal of phenolic compounds from olive mill wastewater. Water Res. 46 (2012),
pp. 789–798.
Blanco, J., Malato, S., Fernández, P., Vidal, A., Morales, A, Trincado, P., de Oliveira, J.C., Minero, C.,
Musci, M., Casalle, C., Brunotte, M., Tratzky, S., Dischinger, N., Funken, K.H., Sattler, C., Vincent,
M., Collares-Pereira, M., Mendes, J.F. & Rangel, C.M.: Compound parabolic concentrator technology
development to commercial solar detoxification applications. Sol. Energy 67 (2000), pp. 317–330.
Blanco-Gálvez, J. & Malato-Rodríguez, S.: Solar Detoxification. UNESCO Publishing, France, 2003.
Bousselmi, L., Geissen, S.U. & Schroeder, H.: Textile wastewater tratment and reuse by solar photocatalysis:
Results from a pilot plant in Tunisia. Water. Sci. Technol. 49 (2004), pp. 331–337.
Braslavsky, S.E.: Glossary of terms used in photochemistry. Pure Appl. Chem. 79 (2007), pp. 293–465.
Brown, C.D. & van Beinum, W.: Pesticide transport via sub-surface drains in Europe. Env. Pollut.157 (2009),
pp. 3314–3324.
Bueno, M.J., Agüera, A., Gómez, M.J., Hernando, M.D., García-Reyes, J.F. & Fernández-Alba, A.R. :
Application of liquid chromatography/quadrupole-linear ion trap mass spectrometry and time of flight
mass spectrometry to the determination of pharmaceuticals and related contaminants in wastewater.
Anal.Chem. 79 (2007), pp. 9372–9384.
Calderón-Preciado D., Jiménez-Cartagena C., Matamoros V. & Bayona, J.M.: Screening of 47 organic
micropollutants in agricultural irrigation waters and their soil loading. Water Res. 45 (2011), pp. 221–231.
Cassano, A.E. & Alfano, O.E.: Reaction engineering of suspended solids heterogeneous photocatalytic
reactors. Catal. Today 58 (2000), pp. 167–197.
Cassano, D., Zapata, A., Brunetti, G., Del Moro, G., Di Iaconi, C., Oller, I., Malato, S. & Mascolo,
G.: Comparison of several combined/integrated biological-AOPs setups for the treatment of munici-
pal landfill leachate: Minimization of operating costs and effluent toxicity. Chem. Eng. J. 172 (2011),
pp. 250–257.
Chong, M.N., Jin, B., Chow, C.W.K. & Saint, C.: Recent developments in photocatalytic water treatment
technology: A review. Water Res. 44 (2010), pp. 2997–3027.
Daughton, C.G. & Ternes, T.A.: Pharmaceuticals and personal care products in the environment: Agents of
subtle change? Environ. Health Pers. 107 (1999), pp. 907–938.
Colina-Márquez, J., Machuca-Martínez, F. & Li Puma, G.: Radiation adsorption and optimisation of solar
photocatalytic reactors for environmental applications. Env. Sci. Technol. 44 (2010), pp. 5112–5120.
De, Angelo L. & Black, B.: London smog disaster, England. In: C. J. Cleveland (ed): Encyclopaedia of Earth.
Environmental Information Coalition, National Council for Science and the Environment Washington
D.C., USA, 2006.
Dillert, R., Cassano, A.E., Goslich, R. & Bahnemann, D.: Large scale studies in solar catalytic wastewater
treatment. Cat. Today 54 (1999), pp. 267–282.
Esplugas, S. & Ollis, D.F.: Economic aspects of integrated (chemical + biological) processes for water
treatment. J. Adv. Oxid. Technol. 2 (1997), pp. 197–204.
Fatta-Kassinos, D., Kalavrouziotis, I.K., Koukoulakis, P.H. & Vasquez, M.I.: The risks associated with
wastewater reuse and xenobiotics in the agroecological environment. Sci. Total Environ. 409 (2011),
pp. 3555–3563.
Fenton, H.J.H.: Oxidation of tartaric acid in presence of iron. J. Chem. Soc. 65 (1894), pp. 899–910.
Fernández-Ibáñez, P., Malato, S. & De Las Nieves, F.J.: Relationship between TiO2 particle size and reactor
diameter in solar photodegradation efficiency. Cat. Today 54 (1999), pp. 195–204.
20 S. Malato et al.

Frank, S.N. & Bard, A.J.: Heterogeneous photocatalytic oxidation of cyanide ion in aqueous solutions at
titanium dioxide powder. J. Am. Chem. Soc. 99 (1977), pp. 303–304.
Fujishima, A. & Honda, K.: Electrochemical photolysis of water at a semiconductor electrode. Nature 238
(1972), pp. 37–38.
Fujishima, A., Rao, T.N. & Tryk, D.A.: Titanium dioxide photocatalysis. J. Photochem. Photobio. C:
Photochem. Review 1 (2000), pp. 1–21.
Gari, L.: Arabic treatises on environmental pollution up to the end of the thirteenth century. Environ. Hist. 8
(2002), pp. 475–488.
Gernjak, W., Maldonado, M.I., Malato, S., Cáceres, J., Krutzler, T., Glaser, A. & Bauer, R.: Pilot-plant
treatment of olive mill wastewater (OMW) by solar TiO2 photocatalysis and solar photo-Fenton. Sol.
Energy 77 (2004), pp. 567–572.
Gernjak, W., Fuerhacker, M., Fernández-Ibáñez, P., Blanco, J. & Malato, S.: Solar photo-Fenton treatment—
Process parameters and process control. Appl.Catal. B: Environ. 64 (2006), pp. 121–130.
Gernjak, W., Krutzler, T., Malato, S. & Bauer, R.: Photo-Fenton treatment of olive mill wastewater applying
a combined Fenton/flocculation pretreatment. J. Solar Energy Eng. 129 (2007), pp. 53–59.
Giménez, J., Curcó, D. & Queral, M.A.: Photocatalytic treatment of phenol and 2,4-dichlorophenol in a solar
plant in the way to scaling-up. Cat. Today 54 (1999), pp. 229–244.
Gogate, P.R. & Pandit, A.B.: A review of imperative technologies for wastewater treatment. I: Oxidation
technologies atambient conditions. Adv. Environ. Res. 8 (2004), pp. 501–551.
Gómez, M.J., Gómez-Ramos, M.M., Agüera, A., Mezcua, M., Herrera, S. & Fernández-Alba, A.R.: A new
gas chromatography/mass spectrometry method for the simultaneous analysis of target and non-target
organic contaminants in waters. J. Chrom. A 1216 (2009), pp. 4071–4082.
Goswami, D.Y.: A review of engineering developments of aqueous phase solar photocatalytic detoxification
and disinfection processes. J. Sol. En.Eng. Trans. ASME 119 (1997), pp. 101–107.
Hai, F.I., Yamamoto, K. & Fukushi, K.: Hybrid treatment systems for dye wastewater. Crit. Rev. Env. Sci.
Technol. 37 (2007), pp. 315–377.
Herrmann, J.M.: Photocatalysis fundamentals revisited to avoid several misconceptions. Appl. Cat. B:
Environ. 99 (2010), pp. 461–468.
Hignite, C. & Azarnoff, D.L.: Drugs and drug metabolites as environmental contaminants: Chlorophenoxy-
isobutyrate and salicylic acid in sewage water effluents. Life Sci. 20 (1977), pp. 337–341.
Hogenboom, A.C., van Leerdam, J.A. & de Voogt, P.: Accurate mass screening and identification of emerging
contaminants in environmental samples by liquid chromatography-hybrid linear ion trap Orbitrap mass
spectrometry. J. Chrom. A 1216 (2009), pp. 510–519.
Hong, S., Candelone, J.P., Patterson, C.C. & Boutron, C.F.: History of ancient copper smelting pollution
during roman and medieval times recorded in Greenland ice. Science 272 (1996), pp. 246–249.
Kavitha, V. & Palanivelu, K.: The role of ferrous ion in Fenton and photo-Fenton processes for the degradation
of phenol. Chemosphere 55 (2004), pp. 1235–1243.
Kümmerer, K.: Antibiotics in the aquatic environment – A review – part I. Chemosphere 75 (2009),
pp. 435–441.
Lagarde, F., Tusseau-Vuillemin, M.H., Lessard, P., Hèduit, A., Dutrop, F. & Mouchel, J.M.: Variabil-
ity estimation of urban wastewater biodegradable fractions by respirometry. Water Res. 39 (2005),
pp. 4768–4778.
Likens, G.E., Driscoll, C.T. & Buso, D. C.: Long-term effects of acid rain: Response and recovery of a
forested ecosystem. Science 272 (1996), pp. 244–246.
Lipczynska-Kochany, E. & Kocjany, J.: Effect of humic substances on the Fenton treatment of wastewater at
acidic and neutral pH. Chemosphere 73 (2008), pp. 745–750.
Lucas, M.S., Mosteo, R., Maldonado, M.I., Malato, S. & Peres, J.A.: Solar Photochemical treatment of
winery wastewater in a CPC reactor. J. Agr. Food Chem. 57 (2009), pp. 11,242–11,248.
Malato, S., Blanco, J., Vidal, A. & Richter, C.: Photocatalysis with solar energy at a pilot-plant scale: An
overview. Appl. Catal. B: Environ. 37 (2002), pp. 1–15.
Malato, S., Blanco, J., Maldonado, M.I., Fernandez, P., Alarcon, D., Collares, M., Farinha, J. & Correia, J.:
Engineering of solar photocatalytic collectors. Sol. En. 77 (2004), pp. 513–524.
Malato, S., Fernández-Ibañez, P., Maldonado, M.I., Blanco, J. & Gernjak, W.: Decontamination and
disinfection of water by solar photocatalysis: Recent overview and trends. Cat. Today 147 (2009),
pp. 1–59.
Mantzavinos, D. & Psillakis, E.: Enhancement of biodegradability of industrial wastewaters by chemical
oxidation pre-treatment. J. Chem. Technol. Biotechnol. 79 (2004), pp. 431–454.
Decontamination of water by solar irradiation 21

Martínez-Bueno, M.J., Gómez, M.J., Herrera, S., Hernando, M.D., Agüera, A. & Fernández-Alba, A.R.:
Occurrence and persistence of organic emerging contaminants and priority pollutants in five sewage
treatment plants of Spain: Two years pilot survey monitoring. Env. Poll.164 (2012), pp. 267–273.
Mosteo, R., Ormad, M.P. & Ovelleiro, J.L.: Photo-Fenton processes assisted by solar light used as preliminary
step to biological treatment applied to winery wastewaters. Water Sci. Tech. 56 (2007), pp. 89–94.
Neumann, M., Schulz, R., Schafer, K., Muller, W., Mannheller, W. & Liess, M.: The significance of entry
routes as point and non-point sources of pesticides in small streams. Water Res. 36 (2002), pp. 835–842.
Oller, I., Malato, S., Sánchez-Pérez, J.A., Gernjak, W., Maldonado, M.I., Pérez-Estrada, L.A. & Pulgarín, C.:
A combined solar photocatalytic-biological field system for the mineralization of an industrial pollutant
at pilot scale. Cat. Today 122 (2007), pp. 150–159.
Oller, I., Malato, S., Sánchez-Pérez, J.A., Maldonado, M.I., Gernjak, W., Pérez-Estrada, L.A., Muñoz, J.A.,
Ramos, C. & Pulgarín, C.: Pre-industrial-scale combined solar photo-Fenton and immobilised biomass
activated-sludge bio-treatment. Ind. Eng. Chem. Res. 46 (2007), pp. 7467–7475.
Oller, I., Malato, S. & Sánchez-Pérez, J.A.: Combination of advanced oxidation processes and biological
treatments for wastewater decontamination-a review. Sci. Tot. Environ. 409 (2011), pp. 4141–4166.
Pal, A., Yew-Hoong Gin, K., Yu-Chen Lin, A. & Reinhard, M.: Impacts of emerging organic contaminants
on freshwater resources: Review of recent occurrences, sources, fate and effects. Sci. Tot. Environ. 408
(2010), pp. 6062–6069.
Parilti, N.B. & Akten, D.: Response surface methodological approach and the kinetic study for the assessment
of the photodegradation of a pulp mill effluent with H2 O2 /Fe(III)/solar UV. Fres. Environ. Bull. 20 (2011),
pp. 1301–1400.
Petrovic, M. & Barceló, D.: Liquid chromatography-mass spectrometry in the analysis of emerging
contaminants. Anal. Bioanal. Chem. 385 (2006), pp. 422–424.
Pietrogrande, M.C. & Basaglia, G.: GC-MS analytical methods for the determination of personal-care
products in water matrices. TrAC 26 (2007), pp. 1086–1094.
Pignatello, J., Oliveros, E. & McKay, A.: Advanced oxidation processes for organic contaminant destruction
based on the fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Technol. 36 (2006), pp. 1–84.
Pulgarín, C., Invernizzi, M., Parra, S., Sarria, V., Polaina, R. & Péringer, P.: Strategy for the coupling of
photochemical and biological flow reactions useful in mineralization of biocalcitrant industrial pollutants.
Cat. Today 54 (1999), pp. 341–352.
Rizzo, L.: Bioassays as a tool for evaluating advanced oxidation processes in water and wastewater treatment.
Water Res. 45 (2011), pp. 4311–4340.
Robles-Molina, J., Gilbert-López, B., García-Reyes, J.F. & Molina-Díaz, A.: Determination of organic
pollutants in sewage treatment plant effluents by gas chromatography high-resolution mass spectrometry.
Talanta 82 (2010), pp. 1318–1324.
Rodríguez, E., Márquez, G., Carpintero, J.C., Beltrán, F.J. & Álvarez, P.: Sequential use of bentonites and
solar photocatalysis to treat winery wastewater. J. Agric. Food Chem. 56 (2008), pp. 11956–11961.
Sarria, V., Parra, S., Adler, N., Péringer, P. & Pulgarín, C.: Recent developments in the coupling of pho-
toassisted and aerobic biological processes for the treatment of biorecalcitrant compounds. Cat. Today 76
(2002), pp. 301–315.
Sarria, V., Deront, M., Péringer, P. & Pulgarin, C.: Degradation of a biorecalcitrant dye precursor present in
industrial wastewaters by a new integrated iron (III) photoassisted-biological treatment. Appl. Catal. B:
Environ. 40 (2003) pp. 231–246.
Sarria, V., Kenfack, S., Guillod, O. & Pulgarin, C.: An innovative coupled solar-biological system at field
pilot scale for the treatment of biorecalcitrant pollutants. J. Photochem. Photobiol. A: Chem. 159 (2003),
pp. 89–99.
Sattler, C., Funken, K.-H., de Oliveira, L., Tzschirner, M. & Machado, A.E.H.: Paper mill wastewater
detoxification by solar photocatalysis. Wat. Sci. Technol. 49 (2004), pp. 189–193.
Savin, I.I. & Butnaru, R.: Wastewater characteristics in textile finishing mills. Env. Eng. Man. 7(2008),
pp. 859–864.
Schueler, F.W.: Sex-hormonal action and chemical constitution. Science 103 (1946), pp. 221–223.
Scott, J.P. & Ollis, D.F.: Integration of chemical and biological oxidation processes for water treatment:
Review and recommendations. Env. Prog. 14 (1995), pp. 88–103.
Scott J.P. & Ollis, D.F.: Integration of chemical and biological oxidation processes for water treatment II:
Recent illustrations and experiences. J. Adv. Oxid. Technol. 2 (1997), pp. 374–381.
Seinfeld, J.H. & Pandis, S.N.: Atmospheric chemistry and physics – from air pollution to climate change.
John Wiley and Sons Inc., 1998.
22 S. Malato et al.

Singh, K. & Arora, S.: Removal of synthetic textile dyes from wastewaters: A critical review on present
treatment technologies. Crit. Rev. Env. Sci. Technol. 41 (2011), pp. 807–878.
Sirtori, C., Zapata, A., Oller, I., Gernjak, W., Agüera, A. & Malato, S.: Solar photo-Fenton as finishing step
for biological treatment of a pharmaceutical wastewater. Env. Sci. Tech. 43 (2009), pp. 1185–1191.
Skark, C., Zullei-Seibert, N., Willme, U., Gatzemann, U. & Schlett, C.: Contribution of non-agricultural
pesticides to pesticide load in surface water. Pest. Manag. Sci. 60 (2004), pp. 525–530.
Skocaj, M., Filipic, M., Petkovic, J. & Novak, S.: Titanium dioxide in our everyday life; is it safe? Radiol.
Oncol. 45 (2011), pp. 227–247.
Sluczewski, A. & Roth, P.: Effects of androgenic and estrogenic metamorphoses of amphibians. Obst. Gyn.
47 (1948), pp. 164–176.
Spengler, J.D. & Sexton K.: Indoor air pollution: A public health perspective. Science 221 (1983), pp. 9–17.
Suty, H., De Traversay, C. & Cost, M.: Applications of advanced oxidation processes: Present and future.
Water Sci. Technol. 49 (2004), pp. 227–233.
United States Environmental Protection Agency, 1996. Prevention, pesticides and toxic substances (7101).
Fates, transport and transformation test guidelines OPPTS 835.3200 Zahn-Wellens/EMPA Test. EPA
712-C-96- 084.
Vilar, V.J.P., Maldonado, M.I., Oller, I., Malato, S. & Boaventura, R.A.R.: Solar treatment of cork boiling
and bleaching wastewaters in a pilot plant. Water Res. 43 (2009), pp. 4050–4062.
Vilar, V.J.P., Silva, T.F.C.V., Santos, M.A.N., Fonseca, A., Saraiva, I. & Boaventura, R.A.R.: Evaluation of
solar photo-Fenton parameters on the pre-oxidation of leachates from a sanitary landfill. Sol. En. 86
(2012), pp. 3301–3315.
Wiszniowski, J., Robert, D., Surmacz-Gorska, J., Miksch, K., Malato, S. & Weber, J.-V.: Landfill leachate
treatment methods: A review. Environ. Chem. Lett. 4 (2006), pp. 51–61.
World Bank: Environmental, health, and safety guidelines for textile manufacturing, international finance cor-
poration, World Bank group, 2007. Available at: http://www.ifc.org/ifcext/sustainability.nsf/Attachments
ByTitle/gui_EHSGuideline2007_TextilesMfg/$FILE/Final+-+Textiles+Manufacturing.pdf, (accessed 4
October 2012).
Zapata, A., Oller, I., Rizzo, L., Hilgert, S., Maldonado, M.I., Sánchez-Pérez, J.A. & Malato, S.: Evaluation
of operating parameters involved in solar photo-Fenton treatment of wastewater: Interdependence of
initial pollutant concentration, temperature and iron concentration. Appl. Catal. B: Environ. 97 (2010),
pp. 292–298.
Zapata, A., Oller, I., Sirtori, C., Rodríguez, A., Sánchez-Pérez, J.A., López, A., Mezcua, M. & Malato, S.:
Decontamination of industrial wastewater containing pesticides by combining large-scale homogeneous
solar photocatalysis and biological treatment. Chem. Eng. J. 160 (2010), pp. 447–456.
Zhang, W., Zou, L. & Wang, L.: Photocatalytic TiO2 /adsorbent nanocomposites prepared viawet chemical
impregnation for wastewater treatment: A review. Appl. Catal. A: Gen. 371 (2009), pp. 1–9.
CHAPTER 2

Reduction of pentavalent and trivalent arsenic by


TiO2 -photocatalysis: An innovative way of arsenic removal

Marta I. Litter, Ivana K. Levy, Natalia Quici, Martín Mizrahi, Gustavo Ruano,
Guillermo Zampieri & Félix G. Requejo

2.1 INTRODUCTION

Arsenic in drinking water constitutes nowadays a serious problem, affecting the health of several
million people all over the world. Only in the Gangetic delta regions of Bangladesh and West
Bengal in India, it has emerged as an environmental health catastrophe with more than 100 million
people estimated to be at risk (Naidu, 2012); in Latin America, 14 million people could be affected
(Litter et al., 2012). Ingestion of more than 100 mg of the element causes acute poisoning,
but ingestion of small amounts of As for a long period leads to the occurrence of arsenicosis
or Chronic Regional Endemic Hydroarsenicism (hidroarsenicismo crónico regional endémico,
HACRE, in Spanish), responsible for skin alterations and cancer (Bundschuh et al., 2010; 2012;
Figueiredo et al., 2010; Litter et al., 2008; 2010; 2012; Morgada et al., 2008, 2010a). The World
Health Organization (2011) recommends 10 µg L−1 as the maximum allowable As concentration
in drinking water, value taken by most national regulatory agencies. Arsenic pollution in water
can be originated in anthropic activities (mining, use of biocides, wood preservers) but most
pollution is natural, coming from dissolution of minerals in surface or groundwaters, or volcanic
processes (Bundschuh et al., 2010; Litter et al., 2010). Predominant As forms in natural ground-
and surface waters (at neutral pH) are arsenate (As(V), as H2AsO− 2−
4 and HAsO4 ) and arsenite
(As(III), as neutral As(OH)3 ). Methods for As removal from waters are urgent, but they should
take into account that while As(V) can be easily removed by conventional treatments such as ion
exchange or adsorption techniques, As(III) removal is more difficult due to its nonionic form
in aqueous solutions at pH < 9 (Litter et al., 2010). Recently, we focused on the study of the
conversion of As(III) and As(V) to As(0) for immobilization of As dissolved in water by TiO2
photocatalysis (Levy et al., 2012).
Transformation to As(V) makes easier the application of conventional technologies as ion
exchange and adsorption. However, new emerging techniques should be investigated to offer
low-cost solutions to the arsenic problem, especially for low-income populations, as mentioned
before (Litter et al., 2008; 2010; Morgada et al., 2008).
Figure 2.1 presents the Latimer diagram for stable As species, and it shows that reduction
of As(V) into As(III) and even into As(0) is thermodynamically possible with mild reductants;
further reduction to arsine is more difficult. These processes are even harder at basic pH.
Heterogeneous photocatalysis (HP) is one of the most studied advanced oxidation processes
for water and air treatment. In HP, after the incidence of photons of adequate energy on semicon-
ductor particles, electrons in the conduction band (e− +
CB ) and holes in the valence band (hVB ) are
produced, which lead to redox reactions with species present in solution, close to the interface or
adsorbed onto the photocatalyst surface (Hoffmann et al., 1995; Litter, 1999; 2009). The most
used semiconductor in HP is TiO2 , in the commercial Evonik P-25 (P25) form. Equation (2.1)
describes the charge separation in the case of TiO2 :

TiO2 + hν → e− +
CB + hVB (2.1)
23
24 M.I. Litter et al.

Figure 2.1. Latimer diagram for stable As species (all E 0 vs. NHE). Calculations are based on data taken
from Santhanam et al. (1985).

Besides unwanted electron-hole recombination, oxidative and reductive reactions take place
after this primary process; e− +
CB can reduce electron acceptors (A) or hVB can receive an electron
from a donor (D) species, which will suffer oxidation (Eq. 2.2 and 2.3). The probability of these
transfer processes depends on the relative redox positions of the CB and VB of TiO2 in relation
with the redox potential of the neighbor or adsorbed species.
e−
CB + A → A

(2.2)
+
hVB + Dads → D•+
ads (2.3)

When a metal or metalloid ion is present in a HP aqueous system solution, it can undergo
direct reduction or oxidation reactions depending on the reduction potential of the corresponding
couple in relation with the energy levels of e− +
CB and hVB (Litter, 2009).
Oxidative HP reactions of As(III) to As(V) over TiO2 have been thoroughly studied and quite
well understood, including the elucidation of mechanisms and proposed applications to real
systems (Bundschuh et al., 2010; Dutta et al., 2005; Ferguson et al., 2005; Ferguson and Hering,
2006; Fostier et al., 2008; Guan et al., 2012; Jayaweera et al., 2003; Leng et al., 2007; Li et al.,
2009; Litter, 2009; 2010; 2012; Morgada de Boggio et al., 2010b; Nguyen et al., 2008; Park
et al., 2010; Ryu and Choi, 2004; Sharma and Sohn, 2009; Tsimas et al., 2009; Xu and Meng,
2009; Yang et al., 1999; Yoon and Lee, 2005; Yoon et al., 2009; Zhang and Itoh, 2006). In
particular, the mechanism of photocatalytic oxidation of As(III) has been a matter of discussion
for many years, the controversy being centered on determining the major oxidant in the system,
either O− +
2 (Choi et al., 2010; Lee et al., 2002; Ryu et al., 2004; 2006), HO or hVB (Dutta et al.,
2005; Fei et al., 2011; Ferguson et al., 2005; Jayaweera et al., 2003; Xu et al., 2007; Yoon et al.,
2005; 2009). However, whatever the oxidant is, it is not possible to deny the efficiency of the
HP process to transform As(III) into As(V). The real problem is, however, how to remove later
dissolved As(V). Oxidation of As(III) in the presence of O2 is thermodynamically possible, but
kinetically inhibited (Bissen et al., 2001); however, under HP conditions, As(III) oxidation is very
rapid, taking place in time scales of 10–100 min at various concentrations from the micromolar
to the millimolar range. Under laboratory illumination conditions or solar light irradiation, HP
takes place usually through monoelectronic processes, and As(III) should be converted first to
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 25

As(IV) and then to As(V). Actually, oxidation of As(III) to As(IV) is a very favorable process,
which can be easily driven by photoproduced h+ •
VB or by hydroxyl radicals (HO ) produced from
+
water, taking into account the redox level of hVB generated from P25 (≈+2.9 V)1 (Martin et al.,
1994) and the reported (estimated) reduction potential of the As(IV)/As(III) couple (E 0 ≈ +2.4 V)
(Kläning et al., 1989); also formation of As(IV) by reaction with HO•2 /O•−
2 is thermodynamically
feasible:
+
H2 O + hVB → HO• + H+ (2.4)
+
As(III) + hVB → As(IV) (2.5)
• +
As(III) + HO + H → As(IV) + H2 O (2.6)
As(III) + HO•2 /O•−
2 → As(IV) + HO− −
2 /O2 (2.7)

On the contrary, HP reduction of As(V) or As(III) to a lower oxidation state has been scarcely
studied (Ferguson et al., 2005; Jayaweera et al., 2003; Yang et al., 1999), and some authors
indicate that the process might be much slower than the photocatalytic As(III) oxidation (Bissen
et al., 2001). Direct reduction of As(V) by e−
CB in a monoelectronic processes leading to As(IV)
(Eq. 2.8) is a non-thermodynamically possible transformation because of the highly negative
reduction potential of the As(V)/As(IV) couple (E 0 ≈ −1.2 V) (Kläning et al., 1989) compared
to the e−CB reduction level (≈−0.3 V) (Martin et al., 1994):

As(V) + e−
CB → As(IV) (2.8)

However, an indirect reductive mechanism might take place in the presence of sacrificial
electron donors like alcohols (methanol, 2-propanol) or carboxylic acids (formic acid), able
to produce strong reductive radicals. In agreement, Yang et al. (1999), working with TiO2 -
deoxygenated suspensions (pH 3) observed almost negligible As(V) removal under UV-irradiation
in the absence of electron donors, but a nearly complete removal in the presence of methanol
(MeOH) at various concentrations. On the other hand, evidences of As(III) reduction by HP have
been never reported, with the exception of a brief mention in some papers (Ferguson et al., 2005;
Jayaweera et al., 2003). Moreover, there are no references about one-electron reduction potentials
of species formed from As(III) to more reduced states that could shed light to these HP processes.
A very simplified energy diagram regarding the idea of direct HP reduction for As(III) and
indirect HP reduction for As(V) is shown in Scheme 2.1.
A brief mention needs to be done regarding the organoarsenical species. Although inorganic
As(III) and As(V) are the predominant contaminant species in water, removal or transforma-
tion of organoarsenical species (as monomethylarsonic acid and dimethylarsinic acid) cannot
be neglected as they are employed in agriculture and in industrial activities (Xu et al., 2007;
Yoon et al., 2011) and can also be introduced to the environment at some extent through the
As(V) methylation performed by a variety of organisms (Cullen et al., 1989). Also, complex
organoarsenical species (e.g. having substituted phenyl rings in their structure) are used to con-
trol intestinal parasites in poultry and swine (Bednar et al., 2003; Zheng et al., 2010). Some
authors (Xu et al., 2007; 2008; Yoon et al., 2011; Zheng et al., 2010) have studied the use of
photocatalytic systems with TiO2 for the degradation of this type of compounds.
In this chapter, reduction of As(V) and As(III) species by TiO2 -HP under anoxic conditions is
analyzed as a possible method for arsenic removal from water, based on recently published results
(Levy et al., 2012). In spite of this previous original publication, we think that the novelty of our
work merits presenting our results here as a definite way to make less mobile As(V) and As(III)
species present in aqueous media by As(0) formation. Our work is a first step in the study of the
mechanism of As reduction for a further technological application.

1All reduction potentials in this work are standard values vs. NHE.
26 M.I. Litter et al.

Scheme 2.1. Scheme of the reduction of As species over TiO2 under UV light in the presence of MeOH
starting from As(V) (indirect reduction) followed by As(III) direct reduction by e−
CB .

2.2 EXPERIMENTAL SECTION

2.2.1 Materials and methods


All chemicals were of the highest purity and employed as received. TiO2 (Degussa P-25,
P25) was provided by Degussa AG Germany (now Evonik). Arsenic (III) was sodium meta-
arsenite (NaAsO2 , Baker) and arsenic (V) was sodium arsenate dibasic 7-hydrate (Na2 HAsO4 ·
7H2 O, Baker). MeOH (99.9%, for HPLC, Carlo Erba) and silver diethyldithiocarbamate
((C2 H5 )2 NCSSAg, Merck) were used. HClO4 (70–72%, Merck) was used for pH adjustments.
Deionized water (Apema Osmoion, resistivity = 18 M cm) was used for preparation of solutions
and suspensions.
As(V) concentration in solution was measured by spectrophotometry using the arsenomolyb-
date technique (detection limit (DL): 0.01 mg L−1 , Lenoble et al., 2003). Total As in solution was
assayed by a modification of this method developed in our laboratory, as follows. To guarantee
total As oxidation in the 0.025–3.24 mg As(III) L−1 (3.25 × 10−4 –4.2 × 10−2 mM) range, excess
potassium permanganate (0.44 mM) was added to the sample. After 75 min, reagents A and B
of Lenoble’s paper (2003) were added, and spectrophotometric measurements were read only
after 90 min contact time. In some few cases, total As was determined by ICP-OES (Perkin
Elmer Optima 5100 DV, DL: 0.005 mg L−1 , Rasmusen et al., 2010); a difference of only 3%
was observed between this method and that using arsenomolybdate. Quantofix Arsen 10 strips
(Macherey-Nagel) were used for semiquantitative measurements of AsH3 in the headspace of
the photoreactor. This test is a simplified Gutzeit method (Rand et al., 1976a) that transform As
species into AsH3 (DL of this method is 0.01 mg L−1 ). Quantitative determination of AsH3 was
performed by the silver diethyldithiocarbamate spectrophotometric method (AgDDC, Rand et al.,
1976b) adapting the photoreactor with a connection to a U tube containing silver diethyldithiocar-
bamate to allow circulation and bubbling of AsH3 in the colorimetric solution. DL of this method
is between 0.0001 and 0.010 mg L−1 , depending on the aqueous matrix (Rasmusen et al., 2010).
At the end of the HP runs, the solid residues were filtered and carefully dried under N2 atmo-
sphere before analysis. To test As(V) adsorbed onto the solid samples, powders were resuspended
in a 0.8% NaHCO3 solution, previously bubbled with N2 for 10 min, N2 bubbling was kept for
10 min more, and As(V) was analyzed in the resulting solution.
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 27

SEM-EDS analysis was performed by a Fei Company Quanta 200 apparatus. TEM images
were obtained with a TEM-EM 301 Philips apparatus (60 kV). X-ray photoemission spectra
(XPS) were taken with a hemispherical electrostatic energy analyzer (r = 10 cm) using Al Kα
radiation (hν = 1486.6 eV). X-ray absorption near edge spectroscopy (XANES) experiments were
performed using the in-house X-ray absorption spectrometer, Rigaku R-XAS Looper. As K-edge
(11,868 eV) X-ray XANES spectra were measured at room temperature in the transmission mode.
For experiments on HP samples starting from As(V), a fluorescence detector was employed. No
changes in the XANES spectra were observed during the scanning, indicating the stability of the
samples under X-radiation.

2.2.2 Irradiation systems


Most irradiations were performed in a commercial quartz photoreactor well (Photochemical Reac-
tors Ltd.) provided with a medium pressure mercury lamp (125 W, maximum emission at 366 nm)2
surrounded by a thermostatic jacket at 25◦ C acting as IR filter. Actinometric measurements were
made in the cell by the ferrioxalate method in the same conditions as the photocatalytic experi-
ments. A photon flow per unit volume incident on the cell wall (q0n,p /V) of 127 µEinstein s−1 L−1
was calculated.
TiO2 suspensions (180 mL) containing As(V) or As(III) at fixed concentrations without or with
MeOH were bubbled with N2 (0.5 L min−1 ) and magnetically stirred all throughout the reaction
period. The suspensions were initially equilibrated for 30 min in the dark to reach the adsorption
equilibrium of As(III) or As(V) and MeOH onto the TiO2 surface; the decrease of the As concen-
tration was discounted. As expected, no changes in As concentration in solution were observed
in any case under irradiation in the absence of photocatalyst.
As(V) was used at the following initial concentrations: 0.525 mM (40.5 mg As(V) L−1 ),
0.065 mM (5 mg As(V) L−1 ) and 0.013 mM (1 mg As(V) L−1 ), while As(III) was tested at
0.525 mM (40.5 mg As(III) L−1 ) and 0.013 mM (1 mg As(III) L−1 ); pH was initially adjusted
to 3, unless indicated, and was not controlled during the runs. The TiO2 concentration was always
1 g L−1 . MeOH was added at 0.4 M; this concentration was chosen because it was the highest and
optimal concentration used byYang et al. (1999) and because our preliminary experiments showed
no significant differences with MeOH in the 0.4–1.2 M range. Aliquots were withdrawn from the
photoreactor periodically, and filtered through 0.2 µm Millipore membranes before analysis.
Other experiments were performed using a quartz spectrophotometric cell (optical path length
l = 10 mm), not thermostatted, and black light fluorescent compact lamp (Interelec, 20 W). Inci-
dent irradiance was measured with a Spectroline DM-365 XA radiometer and was 700 µW cm−2 .
3 mL of a 0.525 mM As(V) solution (initial pH 7.5, not controlled) were mixed with TiO2 (1 g L−1 )
and MeOH was added at 0.4 M. In these experiments, N2 was only bubbled for 5 min through a
septum before irradiation and then the cell was tightly sealed. At the end of irradiation, samples
were filtered through 0.2 µm Millipore membranes and analyzed.

2.3 RESULTS

2.3.1 As(V) photocatalytic experiments


No changes in As(V) concentration were observed during the irradiation of TiO2 suspensions
containing 0.525 mM As(V), [TiO2 ] = 1 g L−1 and pH 3 in the absence of MeOH (not shown).
However, As(V) decay was dramatic when 0.4 M MeOH was added to the system. In similar
experiments in the dark, a decay of only 3–7% of the As(V) initial concentration was measured
after 30 min. Figure 2.2, shows the fractions of the three arsenic species presents during the
experiment (As(V), As(III) and As(0)) at different irradiation times. At the beginning of the

2 Minor emissions occurred at 245, 254, 265, 280, 302, 313, 408, 436 and 546 nm.
28 M.I. Litter et al.

Figure 2.2. Evolution of the fraction of As species during an irradiation experiment under UV
light over TiO2 in the presence of MeOH starting from As(V). Conditions: [As(V)]0 =
0.525 mM, [MeOH] = 0.4 M, [TiO2 ] = 1 g L−1 , pH 3, experiments under N2 (0.5 L min−1 ),
λmax = 366 nm, q0n,p /V = 127 µEinstein s−1 L−1 , T = 25◦ C.

experiment only As(V) is present and, gradually, As(III) and As(0) are formed. As(III) fraction
was calculated as As(III) = As(total, dissolved) – As(V). Assuming that As(0) was the only other
species formed in high amount (see below for AsH3 production), its evolution was calculated
by mass balance (as As(0) = As(total, dissolved)0 – As(total, dissolved)t ). The fraction of As(V)
decreases with irradiation time while As(III) and As(0) fractions were found to increase steadily.
At around 90 min, when As(V) was almost totally depleted, As(III) began to decay, indicating
that this species can be also photocatalytically transformed, in accordance with results obtained
in the As(III) HP system (see next Section 2.3.2). From 150 min up to the end of the experiment,
a practically total As(V) removal from the solution is reached while As(III) continues to decrease
and As(0) continues to rise. Interestingly, data at 270 min indicate still the presence of 68% of As
in solution, in the form of As(III), the rest being deposited as elemental As.
At the end of the irradiation, a gray-bluish solid layer was found deposited onto the TiO2 surface,
suggesting that As(0) was produced in its most stable allotropic form (α) at room temperature
(Santhanam et al., 1985). A slight increase of pH was observed during the experiments, not higher
than 0.5 units.
Results using Quantofix Arsen 10 strips indicate that, at 270 min, AsH3 amounted to roughly
0.025 mg L−1 (ca. 0.06% of the initial As). A separate experiment was performed up to 160 min
for a quantification of AsH3 by the Gutzeit method; the concentration of AsH3 at that time reached
3.9 × 10−4 mM (0.07% of the initial As).
Similar HP experiments starting from lower As(V) concentrations, i.e., 0.065 and 0.013 mM,
revealed interesting findings. For example, starting from 0.065 and 0.013 mM, an approximately
82% removal was observed after 30 min stirring in the dark (much higher than that obtained starting
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 29

Figure 2.3. Evolution of the fraction of As species during an irradiation experiment under UV light
over TiO2 starting from As(III) (a) in the presence and (b) in the absence of MeOH. Con-
ditions: [As(III)]0 = 0.525 mM, [MeOH] = 0.4 M, [TiO2 ] = 1 g L−1 , pH 3, experiments under
N2 (0.5 L min−1 ), λmax = 366 nm, q0n,p /V = 127 µEinstein s−1 L−1 , T = 25◦ C.

from 0.525 mM, i.e., around 7% maximum adsorption) without darkening of the photocatalyst,
indicative that the decrease was due only to an adsorption process. After 30 min more under
irradiation, As(0) appearance and AsH3 evolution also took place in both cases, while 90% As(V)
removal was achieved in the first case and As remaining in solution was below 10 µg L−1 for
the lowest initial concentration. These results are clearly indicative that reductive pathways took
place under irradiation.
Preliminary experiments performed starting from 0.525 mM As(V) (pH 7.5, [TiO2 ] = 1 g L−1 ,
0.4 M MeOH), using the simplified irradiation system (black light lamp in a quartz cell), with
N2 bubbling only before irradiation, showed a high reduction efficiency: after 30 min under
irradiation, As(V) removal was 65%, with the gray deposit (As(0)) observed on the catalyst.
However, in this case, AsH3 evolution was not detected with the Quantofix® strips, pointing
out to a potential strategy to avoid arsine formation. Experiments to verify these findings are
underway. Similarly to the other experiments detailed in this chapter, there was no decay of As(V)
concentration or darkening of the photocatalyst neither in the dark nor in the absence of MeOH,
indicating that no reduction takes place under these conditions.

2.3.2 As(III) photocatalytic experiments


Experiments starting from As(III) (0.525 mM) in different conditions in the absence or in the
presence of 0.4 M MeOH were performed. Irradiations were done under N2 bubbling at pH 3
(with added HClO4 ) and at pH 10 (no HClO4 added). Figures 2.3a and 2.3b depict the ternary
plots corresponding to the experiments in the presence and in the absence of MeOH at pH 3,
respectively.
For both conditions, at t = 0 only As(III) is present in solution, and over time, As(0) is formed
and gradually accumulated. In the case of Figure 2.3a, in the presence of MeOH, no As(V)
could be observed, while in the absence of the donor (Fig. 2.3b) As(V) was detected in solu-
tion. In the presence of MeOH, no As(III) decay was observed after 30 min stirring in the dark.
After 195 min of irradiation, 38% As(III) was removed. Similar results were obtained with other
MeOH concentrations (0.2 and 0.8 M, not shown). Again, changes of pH were not higher than
30 M.I. Litter et al.

Figure 2.4. Evolution of the fraction of As species during an irradiation experiment under UV light over
TiO2 starting from As(III) (a) in the presence and (b) in the absence of MeOH. Conditions:
[As(III)]0 = 0.525 mM, [MeOH] = 0.4 M, [TiO2 ] = 1 g L−1 , pH 10, experiments under N2
(0.5 L min−1 ), λmax = 366 nm, q0n,p /V = 127 µEinstein s−1 L−1 , T = 25◦ C.

0.5 units. As(V) in solution was not detected. Although As(V) was found in traces in the solid
residue after desorption with NaHCO3 , its quantification in solution was erratic due to its low
concentration.
In contrast with the total lack of reaction of As(V) in the absence of MeOH, a rather important
decay of As(III) (similar to that of total As) was observed under these conditions, although some-
what lower than in the presence of MeOH (27 vs. 38% in 195 min, respectively, cf. Figures 2.3a
and 2.3b. In this case, As(V) was actually detected in solution, the highest amount found corre-
sponded to 0.0021% after 10 min of reaction; consequently As(V) desorbed from the photocatalyst
was detected in trace quantities. These finding are in agreement with the work of Lee and Choi
(2002) that found some As(V) production from As(III) in similar anoxic conditions. A sepa-
rated experiment performed for quantitative measurement (Gutzeit method) of AsH3 formation
indicated at 60 min ca. 0.06% of the initial As transformed into AsH3 , Another experiment was
performed starting from 0.525 mM As(III), 0.4 M MeOH but without HClO4 addition, i.e., at
pH 10. No adsorption was observed in the dark. At this pH, decay of total As was faster than at
pH 3 (47 vs. 38%, respectively, Figs. 2.3a and 2.4a). The most important finding in this system
was the appreciable As(V) production, together with a significant amount of As(0). A high final
As(III) decay (86%) was observed at the end of irradiation. AsH3 , which appeared at 35 min,
reached 0.5 mg L−1 (1.3% of the initial As) at 105 min, being the highest amount produced in all
HP systems. It is important to note that the calculation of As(0) by mass balance is not totally
accurate in this case because AsH3 production was rather important in the last stages. Another
significant remark is that, in contrast with experiments at pH 3, the pH of the solution changed
considerably decreasing from 10 to 8.
The results of similar experiments with 0.525 mM As(III) but in the absence of MeOH indicated
again no adsorption in the dark and a ca. 40% decay of total As, a slightly lower amount than
in the presence of MeOH (Fig. 2.4b). As(V) in solution and As(0) were appreciably produced
with similar formation rates in the first stages. Then, the amount of As(V) in solution increased
very rapidly and was the only species in solution at the end of the experiment, implying the
total As(III) oxidation. AsH3 appeared at 20 min and the maximum concentration reached was
around 0.4 mg L−1 (1.0% of the initial As) at 195 min, in contrast with the same experiment in
the presence of MeOH, where 0.5 mg L−1 were reached at 105 min). 41% of As(0) was formed,
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 31

Figure 2.5. TEM (a) and SEM-EDS (b) results of the analyses of solids obtained from HP experiments
starting from As(III). Conditions: [As(III)]0 = 0.525 mM with MeOH (0.4 M), pH 3; the solid
analyzed by XANES was obtained after a HP experiment in similar conditions but at pH 10.
The TEM analysis of the TiO2 surface has a magnification of 100,000X, scale: 1 mm = 10 nm.

although this value is also affected by error because of the high AsH3 production in the last stages.
The pH decrease was higher than in the presence of MeOH (from 10 to 7).

2.3.3 Analysis of solid residues


As already stated, dark residues, attributed to As(0), were deposited on TiO2 after all HP reactions.
These deposits gradually whitened with time, indicating oxidation to arsenic oxides when exposed
to air. XRD patterns of the solids (not shown) corresponded always to pure P25, without signals of
As species, due probably to the low amount or to the lack of crystallinity of the arsenical deposits.
SEM-EDS and TEM results at the end of a HP experiment performed with 0.525 mM As(III),
0.4 M MeOH and pH 3 are presented in Figure 2.5. The SEM image of the photocatalyst showed
that the deposit was composed of nanoparticles; the EDS analysis revealed that it contained As
in a TiO2 matrix. The TEM image (magnification: 100,000 X, scale: 1 mm = 10 nm) indicates a
particle size between 10 and 15 nm, disposed in chain form. A comparison with pure P25 showed
no appreciable differences, suggesting that the reaction took place without morphological changes
for the photocatalyst.
XAS studies performed at the As K-edge (Lamberti et al., 2003; Wilke et al., 2001) gave
information on the average oxidation states of As. The XANES spectrum shown in Figure 2.6a
corresponds to 0.525 mM As(III), pH 3, and that of Figure 2.6b to 0.525 mM As(V), pH 10 both
figures shows also spectra of an As2 O3 standard. The values of the edge absorption energies for
the sample and As(III) standard were obtained from the maximum in the first derivative of each
spectrum (graph not shown). For the sample coming from the As(III) experiment, the absorption
edge was at 11,867.4 eV, close to the reported value of 11867 eV for As(0) (Bearden and Burr,
1967; Jegadeesan et al., 2010; Quinn et al. 2004). Another characteristic of the spectrum of the
sample obtained after the HP treatment is the weak white line in comparison with the intense
one expected for the oxide (see spectrum of As2 O3 in the same Fig. 2.6). These facts confirm
that As(III) has been reduced to As(0) after the HP treatment. On the other hand, the spectrum
corresponding to samples starting from As(V) (Fig. 2.6b) presents an absorption edge energy of
11,868.7 eV, which corresponds to an intermediate value between those reported for As(0) and
the one for As(III) (11,869.3 eV) (James-Smith et al., 2010). In this case, the final product after
32 M.I. Litter et al.

Figure 2.6. XANES analysis of the solids obtained from HP experiments starting from (a) [As(III)]0 =
0.525 mM, MeOH (0,4 M), pH 3, and (b) [As(V)]0 = 0.525 mM MeOH (0.4 M) pH 10.

the treatment can be assigned to a mixture of As(0) and As(III) and the existence of some As(V)
cannot be ruled out. The full vertical line indicates the position for the reported energy edge
for As(0) species, while the vertical dotted line corresponds to As(III) species calculated as a
maximum at the first derivative spectrum of standard of As(III).
Radiation damage (i.e. oxidation of As(III)) could be discarded since the experiments were
performed in an “in-house” spectrometer with a lower photon flux than that used in a synchrotron
laboratory. Additionally, successive XANES spectra were recorded to verify the stability of the
As state, in contrast with results of Jegadeesan et al. (2010) where the XANES experiments were
performed using synchrotron radiation.
Additionally, Figures 2.7 and 2.8 present XPS spectra in the region of the As 3d peak for the
sample coming from the As(V) experiment (0.525 mM As(V), 0.4 M MeOH, pH 10). The sample
was analyzed before and after one week to evidence the presence of As due to the overlapping
between the As 3d and Ti 3p peaks. For both figures, the spectrum in the panel (a) was taken
approximately 48 h after the end of the experiment, while the spectrum in the panel (b) was taken
after keeping the sample exposed to the air during one week. The spectra are dominated by the Ti
3p and 3s peaks at 37.7 and 62.9 eV, respectively; a satellite of the Ti 3p peak is also clearly visible
at 14 eV larger binding energy (BE). First, the spectrum of the pure TiO2 sample was fitted using
several Voigt functions. Then, this spectrum of TiO2 plus two identical As 3d doublets were used
to fit the other two spectra (each doublet made of two Voigt functions with relative intensities
3:2, separated by 0.69 eV, to represent the spin-orbit split As 3d3/2,5/2 levels; in these fittings, the
spectrum of pure TiO2 as well as the width, position and separation of the two As3d doublets were
kept unchanged, and only the intensities were varied. A Shirley-type background was used in all
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 33

Figure 2.7. XPS analysis of the solid obtained from HP experiments starting from As(V) analyzed
(a) 48 h after the experiments and (b) one week after the experiments. Conditions of HP:
[As(V)]0 = 0.525 mM with MeOH (0.4 M), pH 10.

Figure 2.8. As3d components corresponding to the fittings of the spectra acquired 48 h after insertion of
samples (HP experiment, 0.525 mM As(V), 0.4 M MeOH, pH 3) in the analysis chamber (a)
and after 1 week (b). The dotted lines show the subcomponents assigned to As(0) and As(III),
at low and high BE, respectively.
34 M.I. Litter et al.

the cases. The satellite peaks induced by the two most prominent satellites of the Al Kα X-ray line
were also considered in the fitting. The resulting fits are shown in Figure 2.7.
The spectrum in the panel (a) shows a weak but clear increase of intensity in the region between
42 and 44 eV, where the As 3d contribution is expected for As in different oxidation numbers. In the
spectrum taken after one week, this contribution is more evident and intense in the region around
44 eV. Figure 2.7 shows that the fits of the last two spectra yield different As 3d components. These
components can be observed in more detail in Figure 2.8. The 3d5/2 peak of the low-BE component
is located at 42.1 eV, and corresponds to As(0); the high-BE component is then assigned to As(III).
From this figure, it was concluded that initially there are similar amounts of As(0) and As(III)
(proposed to be formed by oxidation of elemental As in the 48 h previous to the XPS measurement,
process enhanced by contact with TiO2 , as stated by Yang et al. (1999), but, after 1 week in air,
the amount of As(0) was strongly reduced in favor of the As(III) oxidized species.

2.4 DISCUSSION

2.4.1 Mechanisms at acid pH


At pH 3, As(V) is present predominantly as H2AsO− 4 (pK a1 = 2.3; pK a2 = 6.8 and pK a3 = 11.6),
whereas arsenite occurs as As(HO)3 (pK a1 = 9.2; pK a2 = 12.7).
As stated above, As(V) transformation to a more reduced form by a direct reductive photocat-
alytic pathway driven by TiO2 e−CB is not thermodynamically possible, based on the highly negative
standard potential of the As(V)/As(IV) couple (Kläning et al., 1989). However, an indirect reduc-
tive mechanism, without e− CB participation, would be possible in the presence of sacrificial electron
donors, such as MeOH, as found by Yang et al. (1999). Hydroxymethyl radicals, strong reducing
intermediates, are generated by h+ •
BV or HO attack because the energy of the TiO2 -VB allows
easy one-electron oxidation of MeOH (E 0 = 1.45 V, Wang et al., 2004):
+
CH3 OH + hBV → • CH2 OH + H+ (2.9)
• •
CH3 OH + HO → CH2 OH + H2 O (2.10)

In the absence of O2 , the • CH2 OH reductive radical (E0 ≈ −0.90 to −1.18 V, Wardman, 1989)
can donate electrons to the CB (current doubling effect, reaction (2.11)) or be the effective reducing
species of As(V), either in the solid-solution interface or in solution, generating formaldehyde
(reaction (2.12)):

CH2 OH → CH2 O + H+ + e−
CB (2.11)
• +
CH2 OH + As(V) → CH2 O + As(IV) + H (2.12)

Although the reduction potential of • CH2 OH is in the limit for As(V) reduction to As(IV), the
value may be somewhat more negative on the TiO2 surface at pH 3. Experiments not reported yet,
irradiating (λ > 300 nm) a mixture of 1 M KNO3 (used as HO• producer, Zepp et al., 1987), 0.4 M
MeOH, 0.5 mM As(V), at pH 1.5 under N2 , produced an immediate As(V) depletion, confirming
that • CH2 OH formed in this homogeneous system is the species responsible for As(V) reduction;
a more profound analysis of this system is underway. In the HP system, formaldehyde can be
transformed to formic acid (FA) and finally mineralized to CO2 . FA generates a much stronger
reducing agent, CO•− •−
2 (E (CO2 /CO2 ) ≈ −2.0 V, Wardman, 1989), which is a better current
0

doubling agent than the radical from MeOH and can contribute to the reducing process of As(V).
CH2 O → HCOOH → CO2 (2.13)
+
HCOOH + hVB → CO•−
2 + 2H
+
(2.14)
HCOOH + HO• → CO•−
2 + H3 O
+
(2.15)
CO•−
2 + As(V) → CO2 + As(IV) (2.16)
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 35

As(IV) is easily reduced to As(III) (Kläning et al., 1989; Ryu and Choi, 2006) by CB or trapped
electrons (Moser et al., 1991) (reaction (2.17)), by • CH2 OH (reaction (2.18)) or by the equivalent
reaction with CO•−2 (reaction (2.19)):

As(IV) + e− −
CB /etrapp → As(III) (2.17)
As(IV) + • CH2 OH → As(III) + CH2 O + H+ (2.18)
As(IV) + CO•−
2 → As(III) + CO2 (2.19)

If the amount of organic donor is low or null, As(IV) can be reoxidized to As(V) by reactions
(2.20) and (2.21):
+
As(IV) + hVB → As(V) (2.20)
• +
As(IV) + HO + H → As(V) + H2 O (2.21)

This explains the decrease of the photocatalytic efficiency for As(V) reduction at low MeOH
concentrations found by Yang et al. (1989). At the high MeOH concentrations used (0.4 M),
reactions (2.9) and (2.10) would be preferred over competing As(IV) reoxidation by reactions
(2.20) and (2.21).
As(IV) can also disproportionate into As(III) and As(V), a rather rapid reaction (Dutta et al.,
2005; Kläning et al., 1989) proposed to be enhanced on the TiO2 surface (Xu et al., 2005):
2As(IV) → As(III) + As(V) (2.22)

Once formed, As(III) can be also photocatalytically reduced, as actually shown in Figures 2.3
and 2.4. However, Figure 2.3b indicates that As(III) reduction is also possible in the absence of
MeOH, in contrast with As(V). Assuming one-electron steps, As(III) reduction by e− −
CB , etrapp (or

CH2 OH, if present) would lead first to As(II), an unstable oxidation state previously postulated
in the radiolysis of O2 -free solutions of arsenite (Daniels, 1962; Muller et al., 1972):
As(III) + e− −
CB /etrapp → As(II) (2.23)
As(III) + • CH2 OH → As(II) + CH2 O + H+ (2.24)

The one-electron reduction potential for the As(III)/As(II) couple is not known: although the
value should be such that allow direct attack of e−
CB , it might be very similar to the level of the CB,
and the very small driving force would explain the slow As(III) reduction. The fact that As(III)
can be reduced directly in the absence of electron donors suggests that the reduction potential
of the As(III)/As(II) couple should be below the level of the TiO2 CB, an important result that
has to be confirmed by electrochemical experiments. Kinetic reasons due to structural changes
when passing from one As form to the other (unknown so far) can be also proposed. Successive
one-electron transfer and disproportionation reactions would transform As(II) to other reduced
states until As(0) and AsH3 appear as stable products. As(0) formation has been reported in
radiolytic experiments of concentrated arsenite solutions (Muller et al., 1972).
In the absence of MeOH, As(III) HP reduction is somewhat less efficient (27 vs. 38% with
MeOH, cf. Figs. 2.3b and 2.4b), because the anodic reaction is oxidation of water by h+ VB (Litter,
1999; 2009), a sluggish process leading to final oxygen formation:
+
2H2 O + 4hBV → O2 + 4H+ (2.25)

Competition of As(II) for the charge carriers would lead to an unproductive short-circuiting
and reoxidation to As(III):
+
As(II) + hBV → As(III) (2.26)
• +
As(II) + HO + H → As(III) + H2 O (2.27)
36 M.I. Litter et al.

In addition, in the absence of MeOH, and in contrast with the same reaction in the presence
of the donor, As(III) can be not only reduced through reaction (2.23) but also easily oxidized to
As(IV) through reactions (2.5) and (2.6), because it is an effective HO• (k = 9 × 109 M−1 s−1 ) and
hole scavenger (Ryu and Choi, 2006), helped by traces of O2 produced by reaction (2.25). Thus,
As(V) will be ultimately produced through reactions (2.20), (2.21) and (2.22) or by injection to
the CB according to:
As(IV) → As(V) + e−
CB (2.28)

This explains why As(V) could be measured (at a low extent) in the As(III) HP system in the
absence of MeOH (Fig. 2.3b), a process not taking place appreciably in the presence of the donor
(Fig. 2.3a), although traces of As(V) were found adsorbed onto the photocatalyst.
However, the low difference in yield for As(III) HP reduction with MeOH at different con-
centrations and without sacrificial donor can be justified because the alcohols are not the most
suitable electron donors, as indicated in other HP reductions (Prairie et al., 1992). Experiments
with other donors are underway. In all cases, the working conditions will be rather reducing, due
to a combination of processes driven by the species present in the system. For example, AsH3 is
known to be a powerful reductant (Santhanam and Sundaresan, 1985) and As(IV) is also a rather
good reducing species (Kläning et al., 1989).
Another possible cathodic reaction is reduction of H+ (free or adsorbed onto the surface) by e−
CB
to atomic hydrogen (H• ), with final H2 formation; however, this process is thermodynamically
and kinetically restricted on bare P25 (the redox potential for the couple H+ /H• is E0 = −2.3 V
(Breitenkamp et al., 1976), although this value is probably less negative on the TiO2 surface.
However, H+ reduction can be facilitated at acid pH on the TiO2 surface by a metal species (M)
able to build up a TiO2 -heterojunction and to accumulate and store electrons, decreasing at the
same time the recombination of electron-hole pairs. In this way, H• and H2 can be produced,
processes enhanced in the presence of MeOH (Behar and Rabani, 2006):

H+ (M) + e− •
CB → H (M) (2.29)

2H (M) → H2 (2.30)
• •
H (M) + CH3 OH → M + H2 + CH2 OH (2.31)

It can be proposed that As(0), once deposited as small spots of fine nanosized particles,
originates this heterojunction, taking into account the value of the As work function of 5.1 eV
(Allred and Hensley Jr., 1961) a value that can be higher for nanoparticles (Wood, 1981) and
found to be 4.2 eV for P25 (Arango and Carter, 1999; Imanishi et al., 2007), with the formation
of a Schottky barrier of ca. 0.9 eV at the As/TiO2 interface. In this way, electrons can easily flow
from the TiO2 CB to the As(0) nanoparticles, with accumulation and storing of electrons, followed
by H• and H2 formation, thus triggering all subsequent reductive steps. No efforts to measure H2
were made in the present work; however, a low or null concentration can be expected because,
once formed, H• would be rapidly consumed by different routes, forming e.g. AsH3 (Bejan and
Bunce, 2003). Molecular hydrogen is able to reduce directly As(V) to As(0).
In spite of forming a beneficial Schottky barrier, it is also possible that, to a certain extent,
deposited As(0) (which amounts more than 1% on the TiO2 surface at the end of the HP exper-
iments) begins to deactivate the surface by blocking active sites on the photocatalyst. A similar
effect is expected from AsH3 ; actually, both As(0) (Fernandez-Vega et al., 1991) and AsH3 (Quinn
et al., 2004) are known to be poisoning agents in catalytic reactions.

2.4.2 Effect of pH
The global reactions starting from As(V) and leading to As(III), As(0) and AsH3 in the presence
of MeOH at acid pH proceed with increase of pH (reactions (2.32)–(2.34)), while those from
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 37

As(III) (Eqs. 2.35–2.38) proceed without pH changes. In both cases, experimental pH changes
were almost negligible, indicating the complexity of the studied systems.

H2 AsO− +
4 + CH3 OH + H  As(OH)3 + H2 O + CH2 O (2.32)
2H2 AsO− +
4 + 5CH3 OH + 2H  2As(0) + 8H2 O + 5CH2 O (2.33)
H2 AsO− +
4 + 4CH3 OH + H  AsH3 + 4H2 O + 4CH2 O (2.34)
2As(OH)3 + 3CH3 OH  2As(0) + 6H2 O + 3CH2 O (2.35)
As(OH)3 + 3CH3 OH  AsH3 + 3H2 O + 3CH2 O (2.36)
2As(OH)3  2As(0) + 3H2 O + 3/2O2 (2.37)
As(OH)3  AsH3 + 3/2O2 (2.38)

Similar reactions for As(III) in alkaline conditions are:

2H2 AsO− −
3 + 3CH3 OH  2As(0) + 2OH + 3CH2 O + 4H2 O (2.39)
H2 AsO−
3 + 3CH3 OH  AsH3 + 3CH2 O + 2H2 O + OH −
(2.40)
2H2 AsO− −
3  2As(0) + 3/2O2 + 2OH + H2 O (2.41)
H2 AsO−
3 + H2 O  AsH3 + 3/2O2 + OH

(2.42)

At pH 10, the main forms of As(V) and As(III) are respectively HAsO2− 4 and H2AsO3 . The
enhancement of As(III) decay at pH 10 both in the presence and in the absence of MeOH
(Fig. 2.4a,b) and higher As(0) and AsH3 yields obtained at this pH compared to those at pH 3,
can be explained mainly in terms of: (i) the level of the CB that is shifted in a Nernstian way
to a more negative position, making more favorable As(III) reduction by the direct route (if the
As(III)/As(II) couple does not depend on pH); (ii) at pH 10, MeOH is mostly in the anionic form
(CH3 O− ), and the radical formed from this species, • CH2 O− (pK a = 10.7, Laroff and Fessenden,
1973), is a more powerful reductant (E 0 ≈ −1.60–1.81 V, Wardman, 1989).
Another result of working at pH 10 that supports the proposed mechanisms is the higher release
of As(V) in solution. As pH is high, both As(V) and TiO2 are in anionic forms and adsorption is
negligible, once formed, As(V) is desorbed and transferred to the solution. The effect is particularly
important in the absence of MeOH because As(V) will be produced by an immediate capture of
h+ • •−
VB or HO (partly as O , pK a = 11.9, Simic et al., 1973) by As(IV) through reactions (2.20)
and (2.21), favored by the alkaline medium and a higher O2 production through reaction (2.43) by
disproportionation (reaction (2.22)) or by injection of electrons to the CB through reaction (2.28).
+
4HO− + 4hBV → O2 + 2H2 O (2.43)

Concomitantly, As(V) formation by As(III) oxidation consumes hydroxyls (see reactions (2.44)
and (2.45); a decrease of pH was actually observed in both alkaline systems, more important in
the absence of MeOH (2 units in the presence of MeOH and 3 units in its absence).

2H2 AsO− −
3 + O2 + 2OH  2HAsO4 + 2H2 O
2−
(2.44)
H2 AsO−
3 + CH2 O + OH −
 HAsO2−
4 + CH3 OH (2.45)

The deceleration of As(V) production in solution at around 150 min in the presence of MeOH
(Fig. 2.4a) can be ascribed to reduction to As(III) by • CH2 O− , with a continuous cycle of
As(V)/As(III) production. However, this decay is also visible in the absence of the alcohol
(Fig. 2.4b), but here it can be ascribed to depletion of As(III) from which As(V) is rapidly
formed through reactions (2.5), (2.6), (2.20) and (2.21), with As(V) being the only final species
in solution at the end of the run.
38 M.I. Litter et al.

As observed, the photocatalytic As(III) decay is enhanced by increasing pH; however, in this
case, As(III) is not only reduced but also oxidized to As(V), especially in the absence of the organic
donor. Although the processes at pH 10 are not suitable for removing As from the solution, they
shed light, without any doubt, to the HP mechanisms taking place in these complicated systems.

2.4.3 Comparison with previous results


Some previous evidence for the HP reduction of As(III) in anoxic conditions can be extracted by
carefully analyzing preceding results of other authors. Most of these papers, with some excep-
tions, never proposed As(III) reduction, and generally only oxidation to As(V) in solution was
evaluated. The low but observable As(V) formation in HP-As(III) systems in anoxic media was
generally associated to the action of residual O2 , adsorbed or present in the crystalline lattice
of the photocatalyst (Jayaweera et al., 2003; Lee and Choi, 2002; Pena et al., 2005; Ryu and
Choi, 2004; 2006; Xu et al., 2005; Yoon and Lee, 2005; Yoon et al., 2009). When a better elec-
tron acceptor than As(III) was added to the anoxic photocatalytic systems, such as CCl4 , POM,
bromate, Cu2+ or Ag+ , the described result was the enhancement of As(III) oxidation (Ferguson
et al., 2005; Ryu and Choi, 2004; 2006; Xu et al., 2005; Yoon and Lee, 2005; Yoon et al., 2009).
This is actually observed here at pH 10, especially in the absence of MeOH.
Another conclusive evidence of As(III) reduction, observed by some authors but never consid-
ered, was the change in color of the white TiO2 suspension to gray, erroneously attributed to the
accumulation of trapped electrons in TiO2 particles (Lee and Choi, 2002; Pena et al., 2005) with
the exception of Yang et al. (1999) who, similarly to the present work, proposed As(0) formation;
however, the XPS spectrum presented in that paper is not clear and not convincing. The observed
decrease of the photocurrent generated after spiking As(III) in a N2 -saturated solution where a
TiO2 /ITO electrode was immersed might be another evidence of As(III) HP reduction (Ryu and
Choi, 2006).

2.5 CONCLUSIONS

Reduction of As(V) and As(III) species by TiO2 -HP under anoxic conditions can be an innovative
way for arsenic removal from water. As(III) (0.525 mM) can be photocatalytically reduced at pH 3
in the presence and in the absence of MeOH, while HP reduction of As(V) takes place only in the
presence of MeOH. As(0) and AsH3 are the products of this reduction.
The reductive HP process at pH 3 is very efficient for As (V or III) removal at the lowest concen-
trations (e.g., 1 mg L−1 , a common value found in natural polluted groundwaters) because it allows
attaining As levels lower than those established by regulations in drinking water (<10 µg L−1 ).
As(0) formation would be a definite way of immobilizing As(V) and As(III) species present in
aqueous media. However, it is important to remark that extremely toxic AsH3 is formed in these
very anoxic HP systems. Known as a very toxic compound, its possible evolution to the air has to
be cautioned, and use of a fume hood is mandatory. Some strategies to remove or contain AsH3
are envisaged: (i) adsorption on suitable catalysts as those developed elsewhere (Quinn et al.,
2006; Seredych et al., 2010); (ii) treatment by a second TiO2 photocatalytic oxidative step in the
gas phase using supported TiO2 ; (iii) optimization of a simpler and less reducing system, where
nitrogen is bubbled only before irradiation. In addition, other electron donors, such as citric acid
or ethanol, less toxic than MeOH, can be proposed for As(V) removal or to accelerate As(III)
reduction. This will be the subject of future research.
Regarding the applicability of the treatment to reduce As levels in drinking water, it is not
possible to judge at the present stage of knowledge its feasibility. Experiments with natural
waters (at circumneutral pH values), containing possible inhibitors such as dissolved organic
matter (DOM), bicarbonate/carbonate ions, etc., which could affect the chemistry of the redox
reactions in solution and at the water/photocatalyst interface should be taken into account. DOM
can act as a UV filter, decreasing the efficiency of the method.
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 39

ACKNOWLEDGMENTS

This work was performed as part of Agencia Nacional de Promoción Científica y Tecnológica
PICT-512, PICT-2008-00038, PAE-PME-2007-00039, PICT 33432 and CYTED 406RT0282
Thematic Network.

REFERENCES

Allred, A.L. & Hensley, A.L. Jr.: Electronegativities of nitrogen, phosphorous, arsenic, antimony and bismuth.
J. Inorg. Nucl. Chem. 17 (1961), pp. 43–54.
Arango, A.C. & Carter, S.A.: Charge transfer in photovoltaics consisting of interpenetrating networks of
conjugated polymer and TiO2 particles. Appl. Phys. Lett. 74 (1999), pp. 1698–1700.
Bearden, J.A. & Burr, A.F.: Reevaluation of X-ray atomic energy levels. Rev. Mod. Phys. 39 (1967),
pp. 25–142.
Behar, D. & Rabani, J.: Kinetics of hydrogen production upon reduction of aqueous TiO2 nanoparticles
catalyzed by Pd0 , Pt0 , or Au0 coatings and an unusual hydrogen abstraction; steady state and pulse
radiolysis study. J. Phys. Chem. B 110 (2006), pp. 8750–8755.
Bejan, D. & Bunce, N.J.: Electrochemical reduction of As(III) and As(V) in acidic and basic solutions. J. Appl.
Electrochem. 33 (2003), pp. 483–489.
Bissen, M., Vieillard-Baron, M.M., Schindelin, A.J. & Frimmel, F.H.: TiO2 -catalyzed photooxidation of
arsenite to arsenate in aqueous samples. Chemosphere 44 (2001), pp. 751–757.
Breitenkamp, M., Henglein, A. & Lilie, J.: Mechanism of the reduction of lead ions in aqueous solution
(a pulse radiolysis study). Ber. Bunsen-Ges. Phys. Chem. 80 (1976), pp. 973–979.
Bundschuh, J., Litter, M., Ciminelli, V., Morgada, M.E., Cornejo, L., GarridoHoyos, S., Hoinkis, J.,
Alarcón-Herrera, M.T., Armienta, M.A. & Bhattacharya, P.: Emerging mitigation needs and sustain-
able options for solving the arsenic problems of rural and isolated urban areas in Iberoamerica – A critical
analysis. Water Res. 44 (2010), pp. 5828–5845.
Bundschuh, J., Litter, M.I., Parvez, F., Román-Ross, G., Nicolli, H.B., Jean, J.-S., Liu, C.-W., López D.,
Armienta, M.A., Gomez Cuevas, A., Cornejo, L., Cumbal, L., Guilherme, L.R.G. & Toujaguez, R.: One
century of arsenic exposure in Latin America: A review of history and occurrence from 14 countries. Sci.
Total Environ. 429 (2012), pp. 2–35.
Choi, W., Yeo, J., Ryu, J., Tachikawa, T. & Majima, T. Environ. Sci. Technol. 44 (2010), pp. 9099–9104.
Cullen, W.R., Reimer, K.J.: Arsenic speciation in the environment. Chem. Rev. 173 (1989),
pp. 713–764.
Daniels, M.: The radiation chemistry of arsenite. Part II. Oxygen-free solution. J. Phys. Chem. (1962),
pp. 1475–1477.
Dutta, P.K., Pehkonen, S.O., Sharma, V.K. & Ray, A.K.: Photocatalytic oxidation of arsenic (III): Eevidence
of hydroxyl radicals. Environ. Sci. Technol. 39 (2005), pp. 1827–1834.
Fei, H., Leng, W., Li, X., Cheng, X., Xu, Y., Zhang, J., Cao, C.: Photocatalytic oxidationof arsenite over
TiO2 : Is superoxide the main oxidant in normal air-saturated aqueous solutions? Environ. Sci. Technol.
45 (2011), pp. 4532-4539.
Ferguson, M.A., Hoffmann, M.R. & Hering, J.G.: TiO2 -photocatalyzed As(III) oxidation in aqueous
suspensions: Reaction kinetics and effects of adsorption. Environ. Sci. Technol. 39 (2005), pp. 1880–1886.
Ferguson, M.A. & Hering, J.G.: TiO2 -photocatalyzed As(III) oxidation in a fixed-bed, flow-through reactor.
Environ. Sci. Technol. 40 (2006), pp. 4261–4267.
Fernandez-Vega, A., Feliu, J.M. & Aldaz, A.: Heterogeneous electrocatalysis on well-defined platinum
surfaces modified by controlled amounts of irreversibly adsorbed adatoms. J. Electroanal. Chem. 305
(1991), pp. 229–240.
Figueiredo, B.R., Litter, M.I., Silva, C.R., Mañay, N., Londono, S.C., Rojas, A.M., Garzón, C., Tosiani, T.,
Di Giulio, G.M., De Capitani, E.M., dos Anjos, J.Â.S.A., Angélica, R.S., Morita, M.C., Paolielo, M.M.B.,
Cunha, F.G., Sakuma, A.M. & Licht, O.A.: Medical geology studies in South America”. In: O. Selinus,
R.B. Finkelman & J.A Centeno (eds): Medical Geology: A Regional Synthesis. Book Series International
Year of Planet Earth. Springer, The Netherlands, 2010, pp. 79–106.
Fostier, A.H., Silva Pereira, M.S., Rath, S. & Guimaraẽs, J.R.: Arsenic removal from water employ-
ing heterogeneous photocatalysis with TiO2 immobilized in PET bottles. Chemosphere 72 (2008),
pp. 319–324.
40 M.I. Litter et al.

Guan, X., Du, J., Meng, X., Sun, Y., Sun, B., Hu, Q. Application of titanium dioxide in arsenic removal from
water: A review. J. Hazard. Mater. 215–216 (2012), pp. 1–16.
Hoffmann, M.R., Martin, S.T., Choi, W. & Bahnemann, D.W.: Environmental applications of semiconductor
photocatalysis. Chem. Rev. 95 (1995), pp. 69–96.
Imanishi, A., Tsuji, E. & Nakato, Y.: Dependence of the work function of TiO2 (rutile) on crystal faces,
studied by a scanning Auger microprobe. J. Phys. Chem. C 111 (2007), pp. 2128–2132.
James-Smith, J., Cauzid, J., Testemale, D., Liu, W., Hazeman, J., Proux, O., Etschmann, B., Philippot, P.,
Banks, D., Williams, P. & Brugger, J.: Arsenic speciation in fluid inclusions using micro-beam X-ray
absorption spectroscopy. Amer. Mineral. 95 (2010), pp. 921–932.
Jayaweera, P.M., Godakumbra, P.I. & Pathiartne, K.A.S.: Photocatalytic oxidation of As(III) to As(V) in
aqueous solutions: A low cost pre-oxidative treatment for total removal of arsenic from water. Curr. Sci.
India 84 (2003), pp. 541–543.
Jegadeesan, G., Al-Abed, S. R., Sundaram, V., Choi, H., Scheckel, K.G. & Dionysiou, D.D.: Arsenic sorption
on TiO2 nanoparticles: Size and crystallinity effects. Water Res. 44 (2010), pp. 965–973.
Kläning, U.K., Bielski, B.H.J. & Sehesteds, K.: Arsenic (IV). Pulse-radiolysis study. Inorg. Chem. 28 (1989),
pp. 2717–2724.
Lamberti, C., Bordiga, S., Bonino, F., Prestipino, C., Berlier, G., Capello, L., D’Acapito, F., Llabrési
Xamena, F.X. & Zecchina, A.: Determination of the oxidation and coordination state of copper on different
Cu-based catalysts by XANES spectroscopy in situ or in operando conditions. Phys. Chem. Chem. Phys.
5 (2003), pp. 4502–4509.
Laroff, G.P. & Fessenden, R.W.: Equilibrium and kinetics of the acid dissociation of several hydroxyalkyl
radicals. J. Phys. Chem. 77 (1973), pp. 1283–1288.
Lee, H. & Choi, W.: Photocatalytic oxidation of arsenite in TiO2 suspension: Kinetics and mechanisms.
Environ. Sci. Technol. 36 (2002), pp. 3872–3878.
Leng, W.H., Cheng, X.F., Zhang, J.Q. & Cao, C.N.: Comment on “Photocatalytic oxidation of arsenite
on TiO2 : Understanding the controversial oxidation mechanism involving superoxides and the effect of
alternative electron acceptors”. Environ. Sci. Technol. 41 (2007), pp. 6311–6312.
Lenoble, V., Deluchat, V., Serpaud, B. & Bollinger, J.-C.: Arsenite oxidation and arsenate determination by
the molybdene blue method. Talanta 61 (2003), pp. 267–276.
Levy, I.K., Mizrahi, M., Ruano, G., Zampieri, G., Requejo, F. & Litter, M.I.: TiO2 -photocatalytic reduction
of pentavalent and trivalent arsenic: Production of elemental arsenic and arsine. Environ. Sci. Technol.
46 (4) (2012), pp. 2299–2308.
Li, Q., Easter, N.J. & Shang, J.K.: As(III) removal by palladium-modified nitrogen-doped titanium oxide
nanoparticle photocatalyst. Environ. Sci. Technol. 43 (2009), pp. 1534–1539.
Litter, M.I.: Transition metal ions in photocatalytic systems. Appl. Catal. B 23 (1999), pp. 89–114.
Litter, M.I.: Treatment of chromium, mercury, lead, uranium and arsenic in water by heterogeneous
photocatalysis. Adv. Chem. Eng. 36 (2009), pp. 37–67.
Litter, M.I., Fernández, R.G., Cáceres R.E., Cobián, D.G., Cicerone, D. & Cirelli, A.F.: Tecnologías de
bajo costo para el tratamiento de arsénico a pequeña y mediana escala. Ingeniería Sanitaria y Ambiental
100 (2008), pp. 41–50.
Litter, M.I., Morgada, M.E. & Bundschuh, J.: Possible treatments for arsenic removal in Latin American
waters for human consumption. Environ. Poll. 158 (2010), pp. 1105–1118.
Litter, M.I., Alarcón-Herrera, M.T., Arenas, M.J., Armienta, M.A., Avilés, M., Cáceres, R.E., Cipriani, H.N.,
Cornejo, L., Dias, L.E., Fernández Cirelli, A., Farfán, E.M., Garrido, S., Lorenzo, L., Morgada, M.E.,
Olmos-Márquez, M.A. & Pérez-Carrera, A.: Small-scale and household methods to remove arsenic from
water for drinking purposes in Latin America. Sci. Total Environ. 429 (2012), pp. 107–122.
Martin, S.T., Herrmann, H., Choi, W. & Hoffmann, M.R.: Time-resolved microwave conductivity (TRMC)
1. TiO2 photoactivity and size quantization. J. Chem. Soc. Faraday Trans. 90 (1994), pp. 3315–3322.
Morgada, M.E. & Litter, M.I.: Arsenic in the Iberoamerican region.The IBEROARSEN Network and a
possible economic solution for arsenic removal in isolated rural zones. e-Terra 5 (2008).
Morgada, M.E., Levy, K., Mateu, M., López, G., Meichtry, M., Bahnemann, D., Dillert, R., Bhattacharya, P.,
Bundschuh, J. & Litter, M: Low-cost solar technologies for arsenic in drinking water. In: N. Kabay, J.
Bundschuh, B. Hendry, M. Bryjak, K. Yoshizuka, P. Bhattacharya & S. Anaç, (eds): The global arsenic
problem: Challenges for safe water production . CRC Press/Balkema, Taylor & Francis Group, London,
UK, 2010a, pp. 209–218.
Morgada de Boggio, M.E., Levy, I.K., Mateu, M., Bhattacharya, P., Bundschuh, J. & Litter, M.I.: Low-
cost technologies based on heterogeneous photocatalysis and zerovalent iron for arsenic removal in the
Chacopampean plain, Argentina. In: J. Bundschuh, M.A. Armienta, P. Bhattacharya„ J. Matschullat,
Reduction of pentavalent and trivalent arsenic by TiO2 -photocatalysis 41

P. Birkle & A.B. Mukherjee (eds): Natural arsenic in groundwater of Latin America – Occurrence, health
impact and remediation. Balkema Publisher, Lisse, The Netherlands, 2010b, pp. 677–686.
Moser, J., Punchihewa, S., Infelta, P.P. & Gratzel, M.: Surface complexation of colloidal semiconductors
strongly enhances interfacial electron-transfer rates. Langmuir 7 (1991), pp. 3012–3018.
Muller, J.C., Ferradini, C. & Pucheault, J.: Radiolyse à très haute intensité des solutions d’arsenite.
Radiochem. Radioanal. Lett. 10 (1972), pp. 53–58.
Naidu, R.: Bioavailability and bio-accessibility of arsenic for ecological and human health risk assessment:
The geological and health interface. In: J.C. Ng., B.N. Noller, R. Naidu J. Bundschuh & P. Bhattacharya:
Understanding the geological and medical interface of arsenic, Proceedings 4th. International Congress
on Arsenic in the Environment (As2012), Cairns, Australia, 22–27 July 2012, Taylor and Francis Group,
London, UK, 2010, pp. 3–10.
Nguyen, T.V., Vigneswaran, S., Ngo, H.H., Kandasamy, J. & Choi, H. C.: Arsenic removal by photo-catalysis
hybrid system. Separat. Purif. Technol. 61 (2008), pp. 44–50.
Park, Y., Lee, S. H., Kang, S. O. & Choi, W.: Organic dye-sensitized TiO2 for the redox conversion of water
pollutants under visible light. Chem. Commun. 46 (2010), pp. 2477–2479.
Pena, M.E., Korfiatis, G.P., Patel, M., Lippincott, L. & Meng, X.: Adsorption of As(V) and As(III) by
nanocrystalline titanium dioxide. Water Res. 39 (2005), pp. 2327–2337.
Prairie, M.R., Evans, L.R. & Martínez, S.L.: Destruction of organics and removal of heavy metals in water
via TiO2 photocatalysis. Chem. Oxid. 2 (1992), pp. 428–441.
Quinn, R., Mebrahtu, T., Dahl, T.A., Lucrezi, F.A. & Toseland, B.A.: The role of arsine in the deactivation of
methanol synthesis catalysts. Appl. Catal. A 264 (2004), pp. 103–109.
Quinn, R., Dahl, T.A., Diamond, B.W. & Toseland, B.A.: Removal of arsine from synthesis gas using a copper
on carbon adsorbent. Ind. Eng. Chem. Res. 45 (2006), pp. 6272–6278.
Rand, M.C., Greenberg, A.E. & Taras M.J. (eds): Standard methods for the examination of water and
wastewater, 14 ed., American Public Health Association, American Water Works Association, Water
Pollution Control Federation (APHA-AWWA-WPCF), Washington D.C., 1976a, pp. 284–286.
Rand, M.C., Greenberg, A.E. & Taras M.J. (eds). Standard methods for the examination of water and
wastewater, 14 ed., American Public Health Association, American Water Works Association, Water
Pollution Control Federation (APHA-AWWA-WPCF), Washington D.C., 1976b, pp. 283–284.
Rasmussen, L. & Jebjerg Andersen, K.: Environmental health and human exposure assessment, Chap-
ter 2, World Health Organization in http://www.who.int/water_sanitation_health/dwq/arsenicun2.pdf.
(accessed October 26, 2010).
Ryu, J. & Choi, W.: Effects of TiO2 surface modifications on photocatalytic oxidation of arsenite: The role
of superoxides. Environ. Sci. Technol. 38 (2004), pp. 2928–2933.
Ryu, J. & Choi, W.: Photocatalytic oxidation of arsenite on TiO2 : Understanding the controversial oxidation
mechanism involving superoxides and the effect of alternative electron acceptors. Environ. Sci. Technol.
40 (2006), pp. 7034–7039.
Santhanam, K.S.V. & Sundaresan, N.S.: Arsenic. In: A.J. Bard & R. Parsons (eds): Standard potentials in
aqueous solutions. Marcel Dekker, New York, USA, 1985, pp. 162–172.
Seredych, M., Mahle, J., Peterson, G. & Bandosz, T.J.: Interactions of arsine with nanoporous carbons: Role
of heteroatoms in the oxidation process at ambient conditions. J. Phys. Chem. C 11 (2010), pp. 6527–6533.
Sharma, V.K. & Sohn, M.: Aquatic arsenic: Toxicity, speciation, transformations, and remediation. Environ.
Int. 35 (2009), pp. 743–759.
Simic, M., Neta, P. & Hayon, E.: Reactions of hydroxyl radicals with unsaturated aliphatic alcohols in
aqueous solution. A spectroscopic and electron spin resonance radiolysis study. J. Phys. Chem. 77 (1973),
pp. 2662–2667.
Tsimas, E.S., Tyrovola, K., Nikolaos P. Xekoukoulotakis, N.P., Nikolaidis, N.P., Diamadopoulos, E. &
Mantzavinos, D.: Simultaneous photocatalytic oxidation of As(III) and humic acid in aqueous TiO2
suspensions. J. Hazard. Mater. 169 (2009), pp. 376–385.
Wagner, C.D., Riggs, W.M., Davis, L.E., Moulder, J.F. & Mullenberg, G.E.: Handbook of X-Ray photoelectron
spectroscopy. Perkin Elmer Corporation, Eden Prairie, MN, USA, 1978.
Wang, C., Pagel, R., Bahnemann, D.W. & Dohrmann, J.K.: Quantum yield of formaldehyde formation in the
presence of colloidal TiO2 -based photocatalysts: Effect of intermittent illumination, platinization, and
deoxygenation. J. Phys. Chem. B 108 (2004), pp. 14,082–14,092.
Wardman, P.J.: Reduction potentials of one-electron couples involving free radicals in aqueous solution.
Phys. Chem. Ref. Data 18 (1989), pp. 1637–1755.
Wood, D.M.: Classical size dependence of the work function of small metallic spheres. Phys. Rev. Lett.
46 (1981), pp. 749.
42 M.I. Litter et al.

Wilke, M., Farges, F., Petit, P.-E., Brown, G.E. Jr. & Martin, F.: Oxidation state and coordination of Fe in
minerals: An Fe K XANES spectroscopic study. Amer. Mineral. 86 (2001), pp. 714–730.
World Health Organization: Arsenic in drinking-water, background document for development of WHO –
Guidelines for drinking water quality. WHO/SDE/WSH/03.04/75/Rev/1, 2011.
Xu, T., Kamat, P.V. & O’Shea, K.E.: Mechanistic evaluation of arsenite oxidation in TiO2 assisted
photocatalysis. J. Phys. Chem. A 109 (2005), pp. 9070–9075.
Xu, T., Cai, Y. & O’Shea, K. E.: Adsorption and photocatalyzed oxidation of methylated arsenic species in
TiO2 suspensions. Environ. Sci. Technol. 41 (2007), pp. 5471–5477.
Xu, Z., Jing, C., Li, F. & Meng, X.: Mechanisms of photocatalytical degradation of monomethylarsonic
and dimethylarsinic acids using nanocrystalline titanium dioxide. Environ. Sci. Technol. 42 (2008),
pp. 2349–2354.
Xu, Z. & Meng, X.: Size effects of nanocrystalline TiO2 on As(V) and As(III) adsorption and As(III)
photooxidation. J. Hazard. Mater. 168 (2009), pp. 747–752.
Yang, H., Lin, W.-Y. & Rajeshwar, K.: Homogeneous and heterogeneous photocatalytic reactions involving
As(III) and As(V) species in aqueous media. J. Photochem. Photobiol. A 123 (1999), pp. 137–143.
Yoon, S. & Lee, J.H.: Oxidation mechanism of As(III) in the UV/TiO2 system: Evidence for a direct hole
oxidation mechanism. Environ. Sci. Technol. 39 (2005), pp. 9695–9701.
Yoon, S., Oh, S.-E., Yang, J.E., Yu, S. & Pak, D.: TiO2 photocatalytic oxidation mechanism of As(III).
Environ. Sci. Technol. 43 (2009), pp. 864–869.
Yoon, S.-H., Lee, S., Kim, T.-H., Lee, M. & Yu, S.: Oxidation of methylated arsenic species by UV/S2 O2− 8 .
Chem. Eng. J. 173 (2011) pp. 290–295.
Zepp, R. G., Hoigné, J. & Bader, H.: Nitrate-induced photooxidation of trace organic chemicals in water.
Environ. Sci. Technol. 21 (1987), pp. 443–450.
Zhang, F.-S. & Itoh, I.: Photocatalytic oxidation and removal of arsenite from water using slag-iron oxide-TiO2
adsorbent. Chemosphere 65 (2006), pp. 125–131.
Zheng, S., Cai, Y. & O’Shea, K.E.: TiO2 photocatalytic degradation of phenylarsonic acid. J. Photochem.
Photobiol. A 210 (2010) pp. 61–68.
CHAPTER 3

Synthesis, characterization and catalytic evaluation of


tungstophosphoric acid immobilized on Y zeolite

Candelaria Leal Marchena, Silvina Gomez, Liliana B. Pierella & Luis R. Pizzio

3.1 INTRODUCTION

The degradation of chemicals present in wastes by heterogeneous photocatalysis is an important


issue and its study is a constantly growing field. An increasing number of papers deal with TiO2
(titania) as one of the most appropriate semiconductor materials to be employed as a photocatalyst,
due to its high activity in the photodegradation of organic compounds, low cost, low toxicity, and
chemical stability (Antonelli et al., 1995; Herrann, 2010; Sakthivel et al., 2004; Sclafani et al.,
1991; Van Grieken et al., 2002; Wang et al., 1998; Yeung et al., 2003; Yu et al., 2000).
It is generally reported that titania performance in the photodegradation of contaminants con-
tained in wastes is influenced by the crystal structure, crystallinity, surface area, porosity, and
band gap energy (Kyoko et al., 1997; Li et al., 1999; Sakulkhaemaruethai et al., 2005), among
other factors.
The fast recombination of the photoinduced electrons and holes can lead to a low photocatalytic
activity. Transition metals or metal oxides were used to avoid the recombination of the electron–
hole pairs of TiO2 -based catalysts (Ikeda et al., 2001; Lee et al., 1992; 1993; Sclafani et al.,
1991), improving the photocatalytic activity.
Additionally, different heteropolyoxometalates (POM) have been added to TiO2 suspensions
(Ozer and Ferry, 2001) or anchored to TiO2 by chemical interactions (Yang et al., 2004), incor-
porated either into TiO2 colloids (Yoon et al., 2001) or into the titania matrix during the TiO2 gel
synthesis (Blanco and Pizzio, 2010; Fuchs et al., 2008; 2009) with the purpose of reducing the
charge recombination in UV-illuminated TiO2 . The capacity of POM as acceptors of the electrons
of UV-irradiated TiO2 suspensions, generated in the conduction band (e− cb ) together with holes in
the valence band (h+ vb ), was demonstrated by Park and Choi (2003) using a photoelectrochemical
method.
Heteropolyoxometalates are widely used as oxidation as well as acid catalysts (Fuchs et al.,
2008; Okuhara et al., 1996; Pizzio et al., 1998). They are also employed as effective homogeneous
photocatalysts in the degradation of organic pollutants in water (Hu et al., 2000; Mylonas and
Papaconstantinou, 1996; Mylonas et al., 1996; Ozer and Ferry, 2000).
POM absorb strongly in the near visible and UV region of the light spectrum (λ < 400 nm). This
absorption corresponds to the ligand-to-metal charge-transfer band and it can generate strongly
oxidizing excited state POM* (reaction (3.1)). They are able to carry out the oxidation of organic
substrates (S) (reaction (3.2)) directly via charge transfer or H-atom abstraction, or indirectly
through the intermediacy of solvent-derived radicals (Ozer and Ferry, 2000). After that, the
corresponding reduced POM is usually reoxidized to their original oxidation state by an electron
acceptor such as dioxygen (reaction (3.3)).

POM + hν → POM∗ (3.1)


∗ − +
POM + S → POM + S (3.2)

POM + O2 → POM + O−
2 (3.3)
43
44 C.L. Marchena et al.

The studies on the photocatalytic behavior of POM have been performed mainly on bulk acids
and their salts. One of the most important drawbacks of POM as catalysts is related to their
high solubility in oxygenated solvents (Okuhara et al., 1996). From a practical point of view,
heterogeneous catalysts are more interesting because their separation from the reaction medium
and reuse are easier. Supporting the POM on solids with high surface area is a useful method
for improving catalytic performance in liquid-solid and gas-solid heterogeneous reactions. It has
been reported that among other factors, the catalytic activity of POM-supported catalyst depends
on: the support type, the concentration and nature of the chemical species present and their degree
of interaction with the support (Blanco and Pizzio, 2010; Okuhara et al., 1996). Silica and NaY
zeolite were employed to prepare promising heterogeneous POM-based photocatalysts (Ozer and
Ferry, 2002). According to Ozer and Ferry (2002) the use of zeolites as support enhances the
“local concentration” of the oxidizable substrate and also the stabilization of charge-transfer state
and transient species such as OH.
Zeolite Y, with uniform pore size and high specific surface area, is considered as good material
to support and stabilize POM (Anandan et al., 2003; Mukai et al., 1997; 2001; Sulikowski et al.,
1996).
In the present work, we attempted to combine the well-known photocatalytic properties of
the tungstophosphoric acid, one of POM most employed in homogeneous photocatalysis, and the
above-mentioned properties of zeolites as POM supports, to synthesized heterogeneous photo-
catalytic materials. These materials can be considered as an alternative to the more widely study
catalysts based on TiO2 .
They were prepared by impregnation of tungstophosphoric acid onto zeolite NH4Y with the
aim to study their influence on the physicochemical, and textural characteristics of the solids,
and to discuss the effect of the preparation variables on the methyl orange (MO) degradation.
To the best of our knowledge, this is the first time that these materials have been successfully
synthesized and tested in the photodegradation of this dye.

3.2 EXPERIMENTAL

3.2.1 Samples preparation


Zeolite NH4Y (Si/Al = 2.47) was provided by Aldrich. The tungstophosphoric acid (TPA) solu-
tions were prepared from H3 PW12 O40 .23H2 O (Fluka p.a.) using distilled water as solvent. The
incorporation of TPA (H3 PW12 O40 ) onto the zeolite matrix was realized by wet impregnation in
a rotary evaporator at 80◦ C. The amount of TPA to be deposited onto the surface of the zeolite
was varied with the purpose of obtaining a TPA concentration of 5, 10, 20, and 30% by weight
in the final solid. They were named NH4YTPA05, NH4YTPA10, NH4YTPA20 and NH4YTPA30,
respectively.
The ammonium salt of tungstophosphoric acid (NH4 )3 PW12 O40 was synthesized by adding
slowly, under vigorous stirring, quantitative amount of 5 M aqueous ammonium carbonate (Merck
p.a.) solution to a 0.1 M aqueous solution of TPA. The excess water was evaporated and the sample
was calcined at 250◦ C in air during 24 h.
The H3 PW12 O40 .23H2 O was treated at the same temperatures during 24 h to obtain the
anhydrous tungstophosphoric acid H3 PW12 O40 .

3.2.2 Sample characterization


3.2.2.1 Textural properties
The specific surface area and the mean pore diameter of the solids were determined from the N2
adsorption-desorption isotherms at the liquid-nitrogen temperature, obtained using Micromeritics
PulseChemisorb 2700 equipment. The solids were previously degassed at 100◦ C for 2 h.
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 45

3.2.2.2 Nuclear magnetic resonance spectroscopy


The 31 P magic angle spinning-nuclear magnetic resonance (31 P MAS-NMR) spectra were recorded
with Bruker Avance II equipment, using the CP/MAS 1 H-31 P technique. A sample holder of 4 mm
diameter and 10 mm in height was employed, using 5 µs pulses, a repetition time of 4 s, and
working at a frequency of 121.496 MHz for 31 P at room temperature. The spin rate was 8 kHz and
several hundred pulse responses were collected. Phosphoric acid 85% was employed as external
reference.

3.2.2.3 Fourier transform infrared spectroscopy


The Fourier transform-infrared (FT-IR) spectra of the solids were obtained using a JASCO 5300
spectrometer and pellets in KBr in the 400–4000 cm−1 wavenumber range.

3.2.2.4 X-Ray diffraction


The X-Ray diffraction (XRD) patterns were recorded with Philips PW-3020 equipment with a
built-in recorder, using CuKα radiation, nickel filter, 20 mA and 40 kV in the high voltage source,
and scanning angle between 5 and 50◦ 2θ at a scanning rate of 2◦ per min.

3.2.2.5 Thermogravimetric analysis and differential scanning calorimetry


The TGA-DSC measurements of the solids were carried out using a Shimadzu DT 50 thermal
analyzer. The thermogravimetric and differential scanning calorimetry analyses were performed
under argon or nitrogen, respectively, using 20–30 mg samples and a heating rate of 10◦ C/min.
The studied temperature range was 20–700◦ C.

3.2.2.6 Diffuse reflectance spectroscopy


The diffuse reflectance spectra (DRS) of the materials were recorded using a UV-visible Lambda
35, Perkin Elmer spectrophotometer, to which a diffuse reflectance chamber Labsphere RSA-PE-
20 with an integrating sphere of 50 mm diameter and internal Spectralon coating is attached, in
the 200-800 nm wavelength range.

3.2.2.7 Potentiometric titration


The acidity of the solids was estimated by means of potentiometric titration. A known mass of
solid was suspended in acetonitrile and stirred for 3 h. Then, the suspension was titrated with
0.05 N n-butylamine in acetonitrile using Metrohm 794 Basic Titrino apparatus with a double
junction electrode.

3.2.3 Photodegradation reaction


The catalytic activity of the materials was evaluated in the photodegradation of methyl orange
dye (Fluka) in water, at 25◦ C. The tests were carried out employing a 125 W high-pressure
mercury lamp (with a maximum emission at about 365 nm) placed inside a Pyrex glass jacket,
thermostatted by water circulation, and immersed in methyl orange dye (MO) solution contained
in a 300 mL cylindrical Pyrex glass reactor. The catalyst is maintained in suspension by stirring,
and air is continuously bubbled. Previously, the MO solution (200 mL, 8.10−5 mol L−1 ) containing
200 mg of catalyst was magnetically stirred in the absence of light for 30 min to ensure that the
adsorption-desorption equilibrium of MO on the surface of the materials is attained. During the
course of the experiments, samples (3 mL) were periodically withdrawn, filtered using a Millipore
syringe adapter (porosity, 0.45 µm) and then analyzed. The variation of the dye concentration as
a function of the reaction time was determined by a UV-visible LAMBDA 35 Perkin Elmer
spectrophotometer of double beam, measuring the absorbance at 465 nm.
In order to evaluate the possibility of TPA leaching during the photocatalytic degradation of
MO, at the end of each experiment, the catalyst was separated by centrifugation, and W was
determined in the liquid phase by atomic absorption spectrometry using a Varian AA Model
46 C.L. Marchena et al.

240 spectrophotometer. The calibration curve method was used, with standards prepared in the
laboratory. The analyses were carried out at a wavelength of 254.9 nm, bandwidth 0.3 nm, lamp
current 15 mA, phototube amplification 800 V, burner height 4 mm and acetylene-nitrous oxide
flame (11:14).

3.3 RESULTS AND DISCUSSION

The specific surface area (S BET ) of the synthesized materials, together with the zeolite NH4Y and
TPA samples, determined from N2 adsorption-desorption isotherms using Brunauer-Emmett-
Teller (BET) method, is listed in Table 3.1.
The N2 adsorption-desorption isotherms of the NH4YTPA05, NH4YTPA10, NH4YTPA20, and
NH4YTPA30 samples exhibit similar characteristics to that of NH4Y zeolite.
According to IUPAC classification, they are Type I isotherms, characteristic of microporous
solids having relatively small external surfaces (Pierella, 2000). The S BET decreases with the
increment of TPA content in the sample.
Taking into account that [PW12 O40 ]3− anions have a diameter of 1.2 nm (Izumi et al., 1983),
the progressive decrease of S BET could be due to the clogging of the pores of NH4Y zeolite.
The XRD patterns of the samples NH4YTPA05, NH4YTPA10, NH4YTPA20 and NH4YTPA30
(Fig. 3.1), show the characteristic peaks of NH4Y zeolite. However, their intensity decreases
in parallel with the increment of TPA content. The XRD diffraction patterns also present an
additional set of peaks, which are different from those characteristic of the tungstophosphoric
acid (H3 PW12 O40 ), its–more common hydrates (H3 PW12 O40 .23H2 O and H3 PW12 O40 .6H2 O), or
other crystalline phases resulting from their transformation (Mioc et al., 1994). These peaks are
similar to those characteristic of the cubic structure of (NH4 )3 PW12 O40 , whose formation takes
place as a result of the interaction between [PW12 O40 ]3− anions and [NH4 ]+ cations present in
the zeolite matrix.
The crystallite size (DC ) of the new crystalline phase, estimated by XRD using the Scherrer
equation, seems to be independent of the TPA content in the prepared materials and they are
similar to those of parent zeolite (Table 3.1).
The interaction between H3 PW12 O40 supports such as SiO2 , TiO2 , or ZrO2 can be assumed to
be of the electrostatic type due to transfer of protons to M-OH according to Lefebvre (1992):
M − OH + H3 PW12 O40 → [M − OH2+ ][H2 PW12 O40 ]

It has been reported that proton transfer from TPA to the amine group, resulting in an elec-
trostatic bond between –NH+ 3 and [H3−x PW12 O40 ] (where 1 < x ≤ 3), is responsible for the
x

efficiently immobilization the heteropolyanion (Pizzio et al., 2002, Hasik et al., 1994).
The 31 P MAS-NMR spectra of NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30
samples (Fig. 3.2) shows two lines of resonance at −15.2 and −13.4 ppm. These lines can be

Table 3.1. Specific surface area (SBET ) and crystallite


size (Dc) of TPA, NH4Y, NH4YTPA samples.

Samples SBET (m2 /g) Dc (nm)

H3 PW12 O40 2 5.0


(NH4 )3 PW12 O40 125 4.8
NH4Y 485 1.1
NH4YTPA05 438 0.9
NH4YTPA10 408 0.9
NH4YTPA20 394 0.9
NH4YTPA30 366 0.9
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 47

Figure 3.1. XRD patterns of H3 PW12 O40 .23H2 O, H3 PW12 O40 , (NH4 )3 PW12 O40 , NH4Y, and the sam-
ples obtained by their impregnation with TPA: NH4YTPA05, NH4YTPA10, NH4YTPA20, and
NH4YTPA30.

Figure 3.2. 31 PMAS-NMR spectra of NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30


samples.

assigned to the presence of the [PW12 O40 ]3− anion and to the dimeric species [P2W21 O71 ]6− ,
respectively (Massart et al., 1977). We can see that in the spectrum of the sample NH4YTPA05,
the line assigned to [P2W21 O71 ]6− is the most intense. However, the signal at −15.2 ppm is the
most important in the rest of the samples. The ratio between the intensity of the signal attributed to
the [PW12 O40 ]3− Keggin anion (δ = −15.2 ppm) and the one assigned to the dimer [P2W21 O71 ]6−
48 C.L. Marchena et al.

Figure 3.3. FT-IR spectra of TPA, NH4Y, NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30
samples.

(δ = −13.4 ppm) increases with the increment of TPA content in the material (NH4YTPA05,
NH4YTPA10, and NH4YTPA20 samples) and then remains constant (sample NH4YTPA30).
The FT-IR spectrum of TPA (Fig. 3.3) shows bands at 1081, 982, 888, 793, 595 and 524 cm−1 ,
which are in accordance with those reported in the literature for the H3 PW12 O40 acid (Massart
et al., 1977). The first five bands are assigned to the stretching vibrations P-Oa , W-Od , W-Ob -
W, W-Oc -W, and to the bending vibration Oa -P-Oa , respectively. The subscripts indicate oxygen
bridging W and the P heteroatom (a), corner-sharing (b) and edge-sharing (c) oxygen belonging
to WO6 octahedra, and terminal oxygen (d).
In the spectra of the NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30 samples
(Fig. 3.3), overlying the characteristic bands of NH4Y zeolite appears a band at 890 cm−1 assigned
to the W-Ob -W stretching, whose relative intensity increases with the increment of TPA in the
materials. Additionally, a broadening of the zeolite band at 790 cm−1 (assigned to the Si-O
stretching vibration (Othman et al., 2006) takes place as a result of its superposition to the
stretching vibration W-Oc -W of TPA.
The main FT-IR bands of the dimer [P2W21 O71 ]6− assigned to the stretching vibrations P-O,
W-O, W-O-W appear at wavenumber values similar to those characteristics of the [PW12 O40 ]3−
anion (Contant, 1987). Taking into account this fact and 31 P MAS-NMR results, we can assume
that the bands at 890 and 790 cm−1 are due to the presence of both [PW12 O40 ]3− and [P2W21 O71 ]6−
species.
According to 31 P MAS-NMR and FT-IR results, the main species present in the samples
(except for NH4YTPA05) is the [PW12 O40 ]3− anion, which was partially transformed into the
[P2W21 O71 ]6− anion during the synthesis and drying steps. This transformation is due to the
limited stability range of the [PW12 O40 ]3− anion in solution, which can be increased by adding
an organic solvent such as ethanol (Kozhevnikov, 1998). At pH 1.5-2, it is reversibly and quickly
transformed into the lacunar species [PW11 O39 ]7− . Pope (1983) has proposed that the following
transformation scheme:
[PW12 O40 ]3− ⇔ [P2 W21 O71 ]6− ⇔ [PW11 O39 ]7−
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 49

Figure 3.4. Thermal analysis diagrams of NH4Y zeolite and H3 PW12 O40 .23H2 O sample.

takes place when the pH is increased. This may be considered as a valid path followed by the TPA
species during the synthesis of the samples.
The DSC diagram of NH4Y zeolite (Fig. 3.4) exhibits an endothermic peak at 136◦ C with a
shoulder at 240◦ C, and another at 306◦ C, due to the desorption of physically adsorbed water, the
decomposition of ammonium ions and structural changes, respectively. They are in agreement
with those reported in the literature (Su et al., 2004).
According to the TGA pattern, the weight loss (25% of the initial mass) associated with the
elimination of water and ammonia, takes place in two main steps below 300o C. The weight loss
ascribed to the structural changes (dehydroxylation) is rather slow (approximately 3%).
The DSC results of H3 PW12 O40 .23H2 O showed two endothermic peaks with maximum at
69 and 185◦ C, and an exothermic one with maximum at 595◦ C. The endothermic features may
be ascribed to the release of physisorbed water and to the dehydration of H3 PW12 O40 .6H2 O
phase, respectively, as in the two weight losses below 250◦ C observed in the TGA diagram. The
exothermic peak may be assigned to the Keggin anion decomposition, which takes place without
appreciable weight loss. Another weight loss, equivalent to 1.5 H2 O per Keggin unit, was observed
in the TG diagram between 250 and 480◦ C. It was assigned to evolution of constitutional water
as a result of deprotonation of the heteropolyacid, as it fits well with the loss of the three protons
of H3 PW12 O40 (Fournier et al., 1992; Okuhara et al., 1996).
During the water loss processes, Keggin anions are not too much disturbed and near 600◦ C
they are decomposed into the constituent oxides or, according to Mioc et al. (1994), transformed
into a new monophosphate bronze type compound PW8 O26 .
Figure 3.5 shows the DSC diagrams of the samples NH4YTPA05, NH4YTPA10, NH4YTPA20,
and NH4YTPA30. They display the characteristic endothermic peaks of NH4Y, although shifted
to slightly lower temperatures with increasing TPA content in the sample.
The DSC diagram of the samples NH4YTPA10, NH4YTPA20, and NH4YTPA30 shows two
exothermic peaks at 558 and 591◦ C, which are assigned to the decomposition of the [P2W21 O71 ]6−
and [PW12 O40 ]3− anions (Ivanov et al., 2004; Okuhara et al., 1996), respectively. In the case of
the sample with a lower TPA content, only the peak assigned to the partial decomposition of the
dimeric species is present.
50 C.L. Marchena et al.

Figure 3.5. DSC profiles of the NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30 samples.

In addition to the above-mentioned peaks, the diagrams show an exothermic peak with maxi-
mum at 582◦ C assigned to the decomposition of the Keggin anion [PW12 O40 ]3− , whose intensity
decreases in parallel with the decrease of TPA content in the samples. The absence of the exother-
mic peak assigned to the [P2W21 O71 ]6− in this set of samples is due, in accordance with 31 P
MAS-NMR, to the low amount of the dimer present in the solids.
The charge-transfer absorption spectra of most nonreduced heteropolyanions obtained by UV-
Vis-DRS appear in the 200–500 nm region and consist of bands that may be ascribed to oxygen-
to-metal transfers. The bulk TPA spectrum presents a band at 212 nm and another broad band that
extends from 250 to 450 nm (Pizzio et al., 1998; Vázquez et al., 2000) assigned to the charge
transfer from bridging or terminal O 2p to W 5d (W-O-W and W-Od , respectively).
The UV-Vis-DRS spectra of the samples NH4YTPA05, NH4YTPA10, NH4YTPA20 and
NH4YTPA30, together with the zeolite NH4Y, are shown in Figure 3.6.
The UV-Vis-DRS spectrum of NH4Y presents two bands at 216 and 300 nm due to charge-
transfer transition of two different Al-O units (Garbowski and Mirodatos, 1982). The first one
that arose from the Al-O charge-transfer transition of four-coordinated framework aluminum,
the second one attributed to structures with highly ordered octahedral symmetry (Zanjanchi and
Razavi, 2001).
The addition ofTPA leads to an increase in the intensity of the band placed at shorter wavelengths
and the appearance of a new one at 263 nm with a shoulder at about 290 nm, which agrees with
the values reported by Fox et al. (Fox et al., 1987) for ammonium salts of the Keggin anion and
the dimeric species, respectively.
The UV-Vis-DRS spectra of the samples display an absorption threshold onset that continuously
shifts to the visible region with the increment of TPA content. The red shift of the absorption
threshold onset is more significant for the NH4YTPA20 and NH4YTPA30 samples.
The band gap energy values (E g ), estimated from UV-Vis-DRS spectra using Kubelka-Munk
remission function (Tandom and Gupta, 1970), are listed in Table 3.2. These values are similar to
those reported in the literature for TiO2 samples (Joselevich and Willner, 1994).
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 51

Figure 3.6. UV-Vis-DRS spectra of the samples NH4Y, NH4YTPA05, NH4YTPA10, NH4YTPA20, and
NH4YTPA30.

Table 3.2. Band gap energy of NH4Y and


NH4YTPA samples.

Samples Eg (eV)

NH4Y 5.10
NH4YTPA05 3.30
NH4YTPA10 3.25
NH4YTPA20 3.20
NH4YTPA30 3.10

The Eg values of NH4Y zeolite without TPA are in agreement with those reported in the literature
(Yan et al., 2008).
The band gap energy values of the samples obtained by impregnation of NH4Y zeolite with
TPA are significantly lower than those of the parent ones. The Eg values decrease slightly with the
increment of the TPA content for the NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30
samples.
The acidity measurements of the catalysts by means of potentiometric titration with n-
butylamine let us estimate the number of acid sites and their acid strength. It was suggested
that the initial electrode potential (Ei ) indicates the maximum acid strength of the sites and the
value of meq amine/g solid where the plateau is reached indicates the total number of acid sites.
The acid strength of these sites may be classified according to the following scale: Ei > 100 mV
(very strong sites), 0 < Ei < 100 mV (strong sites), −100 < Ei < 0 (weak sites) and Ei < −100 mV
(very weak sites) (Pizzio and Blanco, 2003).
The titration curves obtained for the samples NH4YTPA05, NH4YTPA10, NH4YTPA20,
and NH4YTPA30 are shown in Figure 3.7. According to the aforementioned classification,
52 C.L. Marchena et al.

Figure 3.7. Potentiometric titration curves of the NH4YTPA05, NH4YTPA10, NH4YTPA20 and
NH4YTPA30 samples, and (NH4 )3 [PW12 O40 ].

the materials present very strong and strong acid sites, with E i values in the range 220–
290 mV. The Ei value increases with the increment of the TPA content according
to the following order: NH4Y (Ei = 190 mV) < NH4YTPA05 (Ei = 224 mV) < NH4YTPA10
(Ei = 251 mV) < NH4YTPA20 (Ei = 278 mV) < NH4YTPA30 (Ei = 291 mV).
Additionally, the E i values of the samples with higher TPA content are close to that of the
salt (NH4 )3 PW12 O40 (Ei = 288 mV), which is significantly lower than the value reported for
H3 PW12 O40 .21H2 O (Ei = 538 mV) or its partially substituted Cs or K salts (Pizzio and Blanco,
2003). Moreover, the amount of acid sites determined by this technique is practically independent
of the TPA content and it is similar to that of NH4Y zeolite.
Methyl orange (MO), used as a model contaminant, is an azo dye, that exhibit acid-base
equilibrium in solution with pKa = 3.5. Under neutral conditions, MO exists as negative charged
form (Coutinho and Gupta, 2009) with orange color (λmax = 464 nm,  = 24600) (Iida et al.,
2004). The band at 464 nm that is associated with the azo bond (-N≡N-) was used to monitor the
effect of the photocatalysis on the degradation of MO under neutral conditions.
The synthesized materials were tested in the photocatalytic degradation of methyl orange (MO)
at the pH of the suspension obtained by the addition of 200 mg of catalyst to 200 mL of a MO water
solution (C 0MO = 8.10−5 mol L−1 ). The initial and the final pH of the solutions were in the range
6.0–5.0. The MO under these experimental conditions was present in its negative form, and the
surface of the catalysts (point of zero charge in the range 4.8–4.2) was mostly negatively charged.
Therefore, the adsorption of MO on the catalyst surface does not take place in a detectable degree.
The amount of MO degraded after 360 min of irradiation but without catalyst was only of 4%,
and it is mainly due to a photolytic process.
Figure 3.8 shows the evolution of UV–Vis spectra of the solution as a function of irradiation
time. The band at 464 nm decreased during the irradiation, indicating the decomposition of
azo bond and the discoloration of the solution. A new band at wavelength of 249 nm appeared,
suggesting the formation of decomposition intermediates.
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 53

Figure 3.8. Evolution of UV–Vis spectra of a MO water solution as a function of irradiation time.

Figure 3.9. Photocatalytic degradation of MO as a function of the irradiation time.

Figure 3.9 shows the variation of MO concentration (C MO ) as a function of time using the NH4Y,
NH4YTPA05, NH4YTPA10, NH4YTPA20, and NH4YTPA30 samples as catalyst. The diminution
of the MO concentration was only 8% for the NH4Y sample at 300 min under reaction and it
was similar to that achieved without catalyst. However, for the TPA-modified samples, the MO
54 C.L. Marchena et al.

degradation at the same time is significantly higher, and the MO concentration decline (deple-
tion) increases in parallel with the increment of TPA in the NH4YTPA05 (25%) < NH4YTPA10
(53%) < NH4YTPA20 (82%) < NH4YTPA30 (89%). From the comparison of the MO degrada-
tion profiles, we can see that the MO degradation rates increase when the TPA content is raised,
though the increase is lower when the TPA amount changes from 20 to 30%.
The catalytic activity of the bulk (NH4 )3 [PW12 O40 ] salt was measured under the same experi-
mental conditions to be compared with those of the synthesized materials. In this test, the amount
of (NH4 )3 [PW12 O40 ] was fixed in order obtain a TPA amount similar to that contained in the
NH4YTPA30 sample. The MO concentration decreased by 42% at 360 min under reaction and it
was lower than those obtained using the NH4YTPA30. The rather poor catalytic behavior of the
(NH4 )3 [PW12 O40 ] salt could be due to lower specific surface area of ammonium salt (125 m2 g−1 )
in comparison to NH4YTPA30 sample (366 m2 g−1 ). These results are in agreement those reported
by Mizrahi et al. (Mizrahi and Columbus, 2010).
The increment of the catalytic activity when the TPA content increases is principally due to
the direct participation of TPA in the degradation of the organic substrate (reactions (3.1)–(3.33))
and/or in the production of HO• reactive species (reaction (3.4)) that participate in the degradation
of the organic substrate (reaction (3.5)) (Antonaraki et al., 2010):

POM∗ + H2 O → POM− + HO• + H+ (3.4)

HO• + S → oxidation products (3.5)


Additionally, the increment of the catalytic activity can also be due to the lower band gap values
of the samples with higher TPA contents, which would increase the absorption capacity of higher
wavelength radiation.
To evaluate the reusability of the catalysts we chose the NH4YTPA30 catalyst. To this end,
after each photocatalytic experiment, the catalyst was separated from the resulting suspension by
centrifugation, washed with distilled water, dried at 70◦ C and reused.
The percentage of degraded MO after 360 min of irradiation slightly decrease during the first
and the second reuse (85 and 81%, respectively) and then keep constant. The decrease was assigned
to the solubilization of TPA (less than 3% of the TPA content) from the fresh samples during the
first and the second catalytic tests, as was established by atomic absorption spectrometry. Finally,
these results show that after the third use the catalytic performance of the selected samples is
the same, and the synthesized materials can be considered as suitable photocatalysts to degrade
azo dyes.

3.4 CONCLUSIONS

Materials based on tungstophosphoric acid immobilized over NH4Y zeolite were prepared by wet
impregnation.
FT-IR and 31 P MAS-NMR results indicated that the main species present in the samples (except
for NH4YTPA05) is the [PW12 O40 ]3− anion, which was partially transformed into [P2W21 O71 ]6−
anion during the synthesis and drying steps.
The specific surface area of the samples decreased with the increase of the TPA content as a
result of the zeolite pore blocking , attributed to clogging of zeolite pores by the [PW12 O40 ]3−
and [P2W21 O71 ]6− species.
According to DSC and TGA results, the thermal stability of the NH4YTPA materials is similar
to that of their parent zeolite. Additionally, the decomposition of the anions [PW12 O40 ]3− and
[P2W21 O71 ]6− takes place at temperatures similar to those reported for the bulk compounds.
The DRS results indicate that the addition of TPA generates a red shift of the absorption
threshold onset, which is more significant for the samples with the higher TPA content and whose
band gap energy values are similar to those reported for TiO2 .
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 55

According to the results presented in this work, the materials obtained by impregnation of TPA
onto the zeolite matrix, present suitable textural and physicochemical properties to be used as
catalysts in the photocatalytic treatment of wastewater that contains azo dyes.

ACKNOWLEDGEMENTS

The authors thank R. Frenzel, G. Valle and L. Osiglio for their experimental contribution and to
CONICET and UNLP for the financial support.

REFERENCES

Anandan, S., Ryu, S.Y., Cho, W.J. & Yoon, M.: Heteropolytungstic acid (H3 PW12 O40 )-encapsulated into the
titanium exchanged HY (TiHY) zeolite: A novel photocatalyst for photoreduction of methyl orange. J.
Mol. Catal. A: Chem. 195 (2003), pp. 201–208
Antonaraki, S., Triantis, T.M., Papaconstantinou, E. & Hiskia, A.: Photocatalytic degradation of lindane by
polyoxometalates: Intermediates and mechanistic aspects. Catal. Today 151 (2010), pp. 119–124.
Antonelli, D.M., Ying, Y. & Angew, J.: Synthesis of hexagonally-packed mesoporous TiO2 by a modified
sol-gel method. Chem., Int. Ed. Engl. 34 (1995), pp. 2014–2017.
Blanco, M.N. & Pizzio, L.R.: Properties of mesoporous tungstosilicic acid/titania composites prepared by
sol-gel method. Appl. Surf. Sci. 256 (2010), pp. 3546–3553.
Contant Can, R.: Relation entre les tungstophosphates aparentés à l’ánion [PW12 O40 ]3− . Synthèseset
propiétés d’un nouveau polyoxo tungstophosphate lacunaire K10 [P2W20 O70 ].24H2 O. J. Chem. 65 (1987),
pp. 568–573.
Coutinho, C.A. & Gupta, V.K.: Photocatalytic degradation of methyl orange using polymer–titania
microcomposites. J. Colloid Interface Sci. 333:2 (2009), pp. 457–464.
Fournier, M., Feumi-Jantou, C., Rabia, C., Hervé, G. & Launary, S.: Polyoxometalates catalyst materials:
X-ray thermal stability study of phosphorus-containing heteropolyacids H3+x PM12 –xVx O40 · 13–14H2 O
(M = Mo,W; x = 0–1). J. Mat. Chem. 2 (1992), pp. 971–978.
Fox, M.A., Cardona, R. & Gaillard, E.: Photoactivation of metal oxide surfaces: Photocatalyzed oxidation
of alcohols by heteropolytungstates. J. Am. Chem. Soc. 109 (1987), pp. 6347–6354.
Fuchs, V.M., Méndez, L., Blanco, M.N. & Pizzio, L.R.: Mesoporous titania directly modified with
tungstophosphoric acid: Synthesis, characterization and catalytic evaluation. Appl. Catal. A: Gen. 358:1
(2009), pp. 73–78.
Fuchs, V.M., Pizzio, L.R. & Blanco, M.N.: Hybrid materials based on aluminum tungstophosphate or
tungstosilicate as catalysts in anisole acylation. Catal. Today 133–135 (2008), pp. 181–186.
Fuchs, V.M., Soto, E.L., Blanco, M.N. & Pizzio, L.R.: Direct modification with tungstophosphoric acid of
mesoporous titania synthesized by urea-templated sol-gel reactions. J. Colloid Interface Sci. 327 (2008),
pp. 403–411.
Garbowski, E.D. & Mirodatos, C.: Investigation of structural charge transfer in zeolites by UV spectroscopy.
J. Phys. Chem. 86 (1982), pp. 97–102.
Hasik, M., Turek, W., Stochmal, E., Lapkowski, M. & Pron, A.: Conjugated polymer-supported catalysts –
Polyaniline protonated with 12-tungstophosphoric acid. J. Catal. 147:2 (1994), pp. 544–511.
Herrmann, J.M.: Photocatalysis fundamentals revisited to avoid several misconceptions. Appl. Catal B:
Environ. 99 (2010), pp. 461–468.
Hu, C., Yue, B. & Yamase, T.: Photoassisted dehalogenation of organo-chlorine compounds by paratungstate
A in aqueous solutions. Appl. Catal. A: Gen. 194–195 (2000), pp. 99–107.
Iida, Y., Kozuka, T., Tuziuti, T. & Yasui, K.: Sonochemically enhanced adsorption and degradation of methyl
orange with activated aluminas. Ultrasonics 42 (2004), pp. 635–639.
Ikeda, S., Sugiyama, N., Pal, B., Marci, G., Palmisano, L., Noguchi, H., Uosaki, K. & Ohtani, B.: Photo-
catalytic activity of transition-metal-loaded titanium(IV) oxide powders suspended in aqueous solutions:
Correlation with electron-hole recombination kinetics. Phys. Chem. Chem. Phys. 3:2 (2001), pp. 267–273.
Ivanov, A.V., Vasina, T.V., Nissenbaum, V.D., Kustov, L.M., Timfeeva, M.N. & Houzvicka, J.I.: Isomerization
of n-hexane on the Pt-promoted Keggin and Dawson tungstophosphoric heteropoly acids supported on
zirconia. Appl. Catal. A 259:1 (2004), pp. 65–72.
56 C.L. Marchena et al.

Izumi, Y., Matsuo, K. & Urabe, K.: Efficient homogeneous acid catalysis of heteropoly acid and its
characterization through ether cleavage reactions. J. Mol. Catal. 18:3 (1983), pp. 299–314.
Joselevich, E. & Willner, I.: Photosensitization of quantum-size TiO2 particles in water-in-oil microemul-
sions. J. Phys Chem. 98 (1994), pp. 7628–7635.
Kozhevnikov, I.V.: Catalysis by heteropoly acids and multicomponent polyoxometalates in liquid-phase
reactions. Chem. Rev. 98 (1998), pp. 171–198.
Kyoko, K.B., Kazuhiro, S., Hitoshi, K., Kiyomi, O. & Hironori, A.: In-situ FT-IR study on CO2 hydrogenation
over Cu catalysts supported on SiO2 , Al2 O3 , and TiO2 . Appl. Catal. A: Gen. 165 (1997), pp. 391–409.
Lee, W., Gao, Y.M., Dwight, K. & Wold, A.: Preparation and characterization of Titanium(IV) oxide
photocatalysts. Mater. Res. Bull. 27:6 (1992), pp. 685–692.
Lee, W., Do, Y.R., Dwight, K. & Wold, A.: Enhancement of photocatalytic activity of titanium (IV) oxide
with molybdenum (VI) oxide. Mater. Res. Bull. 28:11 (1993), pp. 1127–1134.
Lefebvre, F.: 31 P MAS NMR study of H3 PW12 O40 supported on silica: Formation of (SiOH+ −
2 )(H2 PW12 O40 ).
J. Chem. Soc., Chem. Commun. (1992), p. 756.
Li, X., Cubbage, J.W., Tetzlaff, T.A. & Jenks, W.S.: Photocatalytic degradation of 4-chlorophenol: 1. The
hydroquinone pathway. J. Org. Chem. 64 (1999), pp. 8509–8524.
Massart, R., Contant, R., Fruchart, J., Ciabrini, J. & Fournier M.: 31 P NMR studies on molybdic and
tungstic heteropolyanions. Correlation between structure and chemical shift. Inorg. Chem. 16:11 (1977),
pp. 2916–2921.
Mioc, J.B., Dimitrijevi, R.Z., Davidovic, M., Nedic, Z.P. & Mitrovic, M.M. and Colomban, P.H.: Thermall
induced phase transformations of 12-tungstophosphoric acid hydrate, synthesis and characterization of
PW8 O26 -type bronzes. J. Mater. Sci., 29:14 (1994), pp. 3705–3718.
Mizrahi, D.M. & Columbus, S.S.: Efficient heterogeneous and environmentally friendly degradation of nerve
agents on a tungsten-based POM. J. Hazard. Mater. 179 (2010), pp. 495–499.
Mukai, S.R., Masuda, T., Ogino, I. & Hashimoto, K.: Preparation of encaged heteropoly acid catalyst
by synthesizing 12-molybdophosphoric acid in the supercages of Y-type zeolite. Appl. Catal. A: Gen.
165:1–2 (1997), pp. 219–226.
Mukai, S.R., Lin, L., Masuda, T. & Hashimoto, K.: Key factors for the encapsulation of Keggin-type
heteropoly acids in the supercages of Y-type zeolite. Chem. Eng. Sci. 56:3 (2001), pp. 799–804.
Mylonas, A., Hiskia, A. & Papaconstantinou, E.: Contribution to water purification using polyoxometalates.
Aromatic derivatives, chloracetic acids. J. Mol. Catal. A: Chem. 114 (1996), pp. 191–200.
Mylonas, A. & Papaconstantinou, E.: On the mechanism of photocatalytic degradation of chlorinated phenols
to CO2 and HCl by polyoxometalates. J. Photochem. Photobiol 94:1 (1996), pp. 77–82.
Okuhara, T., Mizuno, N. & Misono, M.: Catalytic chemistry of heteropoly compounds. Adv. Catal. 41 (1996),
pp. 113–252.
Othman, I., Mohamed, R.M. & Ibrahim, I.A.: Synthesis and modification of ZSM-5 with manganese and
lanthanum and their effects on decolorization of indigo carmine dye. Appl. Catal. A: Gen. 299 (2006),
pp. 95–102.
Ozer, R.R. & Ferry, J.L.: Kinetics probes of the mechanism of polyoxometalates-mediated photocatalytic
oxidation of chlorinated organic. J. Phys. Chem. B 104 (2000), pp. 9444–9448.
Ozer, R.R. & Ferry, J.L.: Investigation of the photocatalytic activity of TiO2 -polyoxometalate systems.
Environ. Sci. Technol. 35:15 (2001), pp. 3242–3246.
Ozer, R.R. & Ferry. J.L.: Photocatalytic oxidation of aqueous 1, 2-dichlorobenzene by polyoxometalates
supportedon the NaY zeolite. J. Phys. Chem. B 106 (2002), pp. 4336–4342.
Park, H. & Choi, W.: Photoelectrochemical investigation on electron transfer mediating behaviors of
polyoxometalate in UV-illuminated suspensions of TiO2 and Pt/TiO2 . J. Phys. Chem. B 107 (2003),
pp. 3885–3890.
Pierella, L.B.: Occluded cobalt species over ZSM-5 matrix: Design, preparation, characterization and
magnetic behavior. PhD Thesis. Universidad Tecnológica Nacional, Facultad Regional Córdoba, Córdoba,
Argentina, 2000.
Pizzio, L., Vázquez, P., Kikot, A., Basaldella, E., Cáceres, C. & Blanco, M.: Functionalized SiMCM-41 as
support for heteropolyacid based catyalysts. Stud. Surf. Sci. Catal. 143 (2002), pp. 739–746.
Pizzio, L.R., Cáceres, C.V. & Blanco, M.N.: Acid catalysts prepared by impregnation of tungstophosphoric
acid solutions on different supports. Appl. Catal. A: Gen. 167:2 (1998), pp. 283–294.
Pizzio, L.R. & Blanco, M.N.: Isoamyl acetate production catalyzed by H3 PW12 O40 on their partially
substituted Cs or K salts. Appl. Catal. A: Gen. 255:2 (2003), pp. 265–277.
Pope, M.T.: Heteropoly and isopoly oxometalates. Springer-Verlag, Heidelberg, Germany, 1983.
Synthesis, characterization and catalytic evaluation of tungstophosphoric acid 57

Rocchiccioli-Deltcheff, C., Thouvenot, R. & Franck, R.: Spectres i.r. et Raman d’hétéropolyanions α-
XM12 On− III IV IV V V VI
40 de structure de type Keggin (X=B , Si , Ge , P , As et M=W et Mo ). Spectrochim.
VI

Acta 32A (1976), pp. 587–597.


Sakthivel, S., Shankar, M.V., Palanichamy, M., Arabindoo, B., Bahnemann, D.W. & Murugesan, V.: Enhance-
ment of photocatalytic activity by metal deposition: Characterisation and photonic efficiency of Pt, Au
and Pd deposited on TiO2 catalyst. Water Res. 38:13 (2004), pp. 3001–3008.
Sakulkhaemaruethai, S., Pavasupree, S., Suzuki, Y. & Yoshikawa, S.: Photocatalytic activity of titania
nanocrystals prepared by surfactant-assisted templating method—Effect of calcination conditions. Mater.
Lett. 59:23 (2005), pp. 2965–2968.
Sclafan,A., Mozzanega, M.N. & Pichat, P.J.: Effect of silver deposits on the photocatalytic activity of titanium
dioxide samples for the dehydrogenation or oxidation of 2-propanol. J. Photochem. Photobiol. A: Chem.
59:2 (1991), pp. 181–189.
Sclafani, A., Palmisano, L. & Davì, E.: Photocatalytic degradaton of phenol in aqueous polycrystalline TiO2
dispersions: The influence of Fe3+ , Fe2+ and Ag+ on the reaction rate. J. Photochem. Photobiol. A:Chem.
56:1 (1991), pp. 113–123.
Su, F., Zhao, X.S., Lv, L. & Zhou, Z.: Synthesis and characterization of microporous carbons templated by
NH4 -Y zeolite. Carbon 42 (2004), pp. 2821–2831.
Sulikowski, B., Haber, J., Kubacka, A., Pamin, K., Olejniczak, Z. & Ptaszynski, J.: Novel “ship-in-the-bottle”
type catalyst: Evidence for encapsulation of 12-tungstophosphoric acid in the supercage of synthetic
faujasite. Catal. Lett. 39 (1996), pp. 27–31.
Tandom, S.P. & Gupta, J.P.: Measurement of forbidden energy gap of semiconductors by diffuse reflectance
technique. Phys. Stat. Sol. 38 (1970), pp. 363–367
Van Grieken, R., Aguado, J., López-Muñoz, M.J. & Marugán, J.: Synthesis of size-controlled silica-supported
TiO2 photocatalysts. J. Photochem. Photobiol. A: Chem. 148:1–3 (2002), pp. 315–322.
Vázquez, P., Pizzio, L., Cáceres, C., Blanco, M., Thomas, H., Alesso, E., Finkielsztein, L., Lantaño, B.,
Moltrasio, G. & Aguirre, J.: Silica-supported heteropolyacids as catalysts in alcohol dehydration reactions.
J. Mol. Catal. A: Chem. 161 (2000), pp. 223–232.
Wang, K.H., Tsai, H.H. & Hsieh, Y.H.: The kinetic of photocatalytic degradation of trichloroethylene in gas
phase over TiO2 supported on glass bead. Appl. Catal. B: Environ. 17 (1998) 313–320.
Yan G., Long J., Wang X., Li Z. & Fu X.: Photoactive sites in commercial HZSM-5 zeolite with iron
impurities: An UV Raman study. C.R. Chimie 11: 1–2 (2008) 114–119.
Yang Y., Guo Y., Hu C., Wang Y. & Wang E.: Preparation of surface modifications of mesoporous titania
with monosubstituted Keggin units and their catalytic performance for organochlorine pesticide and dyes
under UV irradiation. Appl. Catal. A: Gen. 273:1–2 (2004), pp. 201–210.
Yeung, K.L., Yau, S.T., Maira, A.J., Coronado, J.M., Soria, J. & Yue, P.L.: The influence of surface properties
on the photocatalytic activity of nanostructured TiO2 . J. Catal. 219:1 (2003), pp. 107–116.
Yoon, M., Chang, J.A, Kim, Y., Choi, J.R., Kim, K. & Lee, S.J.: Heteropoly acid-incorporated TiO2 colloids as
novel photocatalytic systems resembling the photosynthetic. J. Phys. Chem. B 105 (2001), pp. 2539–2545.
Yu, J.C., Lin, J., Lo, D. & Lam, S.K.: Influence of thermal treatment on the adsorption of oxygen and
photocatalytic activity of TiO2 . Langmuir 16 (2000), pp. 7304–7308.
Zanjanchi, M.A. & Razavi, A.: Identification and estimation of extra-framework aluminium in acidic mazzite
by diffuse reflectance spectroscopy. Spectrochimica Acta A 57 (2001), pp. 119–127.
This page intentionally left blank
CHAPTER 4

Kinetic aspects of the photodegradation of phenolic and lactonic


biocides under natural and artificial conditions1

Juan P. Escalada, Adriana Pajares, Mabel Bregliani, Alicia Biasutti, Susana Criado,
Patricia Molina, Walter Massad & Norman A. García

4.1 INTRODUCTION

The continuous and exponential growth of world population is concurrent with higher demands
for more and better food and fiber products. In modern agriculture, application of biocides
has been introduced as one of the tools that favor the achievement of the vital increases and
quality improvement of productions, which are basic necessities for human life and development.
Initially, these chemicals were thought to be only harmful for a target group such as insects,
fungi, nematodes or weed species, but innocuous for humans; however this statement is not
always true. Biocides have proved to be beneficial to control plagues and diseases, but their
residues contaminate soils and natural water reservoirs. For that reason, they constitute a serious
risk for human health and environmental quality.
Significant amounts of agricultural pesticides are usual contaminants of surface waters and
soils (Tomlin, 1994) and their use in crop protection must be conditioned by their persistence in
the environment. Consequently, their thermal, microbiological and photochemical pathways of
degradation are topics of particular research interest.
Within the context of environmentally friendly methods for the elimination of surface water
pollutants, photochemical reactions become a major degradation pathway. In the environment,
photodegradation can take place in two ways: direct, in which a pollutant absorbs energy from the
sunlight (<300 nm); or indirect, in which the substance itself does not absorb the radiation, the
energy is absorbed by an intermediary compound and afterwards transmitted to the target that is
intended to be photodegraded. Here we present a comparative kinetic study between naturally and
artificially promoted photodegradation of known phenolic and lactonic biocides, mainly based on
published data from our group (Escalada et al., 2008; 2011; García, 1994) and interrelated to
findings from other researchers (Feely et al., 1992; Kolar et al., 2003; Konachy et al., 1990;
Mansfield and Richard, 1996; Martiré and González, 2000; Mushtaq et al., 1998).

4.2 PHOTOCHEMICAL DEGRADATION

The photodegradation process depends on the energy supplied by light radiation, which excites the
molecules of the compounds. In nature, pesticides and their degradation products appear finally
as solutions or suspensions in surface waters (Gao et al., 2012). Therefore, the photochemical
pathways of their degradation in aerated solution and under sunlight irradiation, i.e. mimicking
the natural photochemical decay of all these substances, is a topic of great relevance. A suitable
insight that allows a better management of these compounds can be obtained through a systematic
kinetic and mechanistic study of them.

1 Dedicated to the memory of our friend and colleague, Prof. Dr. Francisco Amat-Guerri, whose life was an

example of integrity and commitment to science.

59
60 J.P. Escalada et al.

A compound that absorbs the sunlight can be spontaneously degraded by photochemical pro-
cesses initiated from the electronically excited compound, such as the breakage of molecular
bonds, the direct reaction with ground state oxygen in the medium, and the reaction with reactive
oxygen species (ROS). Examples of ROS are singlet molecular oxygen, O2 (1 g ), superoxide
radical anion (O•−2 ) and hydrogen peroxide (H2 O2 ), generated by energy transfer or electron
transfer from the excited contaminant to oxygen. When the chemical does not absorb any wave-
length of the sunlight or absorbs very little, its degradation in an aquatic natural environment
can also be carried out through its reaction with ROS if another light-absorbing compound—a
photosensitizer— is present in the medium.

4.2.1 Modeling natural photodegradation


The activation of every photochemical process depends on selective light absorption by one of
the compounds involved in it. No chemical, among them biocides, is capable to absorb energy
from the complete electromagnetic spectrum; they only absorb energy at a specific wavelength
range, being transparent to the others. For sunlight-quasi-transparent compounds, an interesting
possibility of photodegradation is through photosensitized processes. The polluted medium is
irradiated with light of wavelengths longer than those absorbed by the compound that has to be
degraded, the contaminant, in the presence of a photosensitizer. In the environment, in surface
natural water bodies, this procedure takes advantage of the presence of traces of some colored
compounds that can be promoted to electronically excited states through absorption of natural
light. The excited forms can directly react with the contaminants, or can produce transient species
able to chemically interact with them. In particular, Riboflavin (Rf), a well-known naturally-
occurring pigment present as traces in water courses, rivers, lakes and seas (Benassi et al., 1967;
Escalada et al., 2006; Momzikoff et al., 1983), has been postulated as a possible sensitizer for the
natural photooxidative degradation of some herbicides and related contaminants (De la Rochette
et al., 2000; Nubbe et al., 1995; Ross and Crosby, 1976; Silber et al., 1976). Upon visible light
irradiation, and in the absence of interacting compounds, Rf generates O2 (1 g ) and O•− 2 with
quantum yields of 0.47 and 0.009, respectively (Chacón, 1988; Krishna et al., 1991).
Although less natural, a second alternative for the decontamination of aquatic systems is the
employment of artificial sensitizers that generate highly reactive transient species. They promote
the photodegradation of the contaminants by known mechanisms.
A sensitizer (S), like Rf or Rose Bengal (RB), absorbs light in a wavelength range—typically
that of the visible light—where the contaminant is transparent. This absorption can produce the
electronically excited states of S; singlet by process 1, 1 S*, and/or triplet by process 2, 3 S*. Both
excited states can be quenched to ground state S through processes 3 and 4, respectively (Fig. 4.1).
An energy transfer process from 3 S* to ground state triplet molecular oxygen, O2 (3 − g ), dis-
solved in the medium, can yield the excited oxygen species O2 (1 g ) – Figure 4.1 (5). O2 (1 g ) can
decay either by collision with surrounding solvent molecules – Figure 4.1(6)– or by interaction
with contaminant and/or S. The last pathway occurs through physical and/or chemical processes,
Figure 4.1(7) and (8), respectively. When 3 S* transfers an electron to the contaminant, it would
give rise to the respective semireduced and semioxidized forms – process (9). Process (10) repre-
sents the generation of the reactive species superoxide radical anion (O•− 2 ), which can chemically
react with the contaminant and/or S, or even with a contaminant radical cation – Figure 4.1(11).

4.2.2 Artificial photodegradation


The photodegradation of a substance as result of direct light absorption without the intervention of
another chemical compound is known as direct photolysis. The process depends on environmental
factors as solar radiation on and below water surface, and intrinsic factors of the chemical itself
like the rate of light absorption and the quantum yield of the photochemical reaction (Fig. 4.2).
The quantum yield is the rate of molecules that react and the number of absorbed photons.
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 61

Figure 4.1. Scheme of main possible processes in the visible-light sensitized photoirradiation under aerobic
conditions of Abamectin (ABA), Bromoxynil (BXN) and Dichlorophen (DCP), all represented
by (B). S: sensitizer (RB or Rf). P: reaction product. O2 (3 −
g ): ground state oxygen. The overall
quenching constant kt is the sum of the rate constants kq and kr . Decay of O− 2 and H2 O2 , by
the influence of superoxide dismutase (SOD) and catalase (CAT) respectively.

Figure 4.2. Scheme of main possible processes in direct photolysis under aerobic conditions of ABA, BXN
and DCP, all represented by B. hν: light; P: products; O2 (3 −
g ): dissolved ground state oxygen.

This parameter is useful to determine the relative weight of the direct photolysis process under
environmental conditions against other degradation processes.
In the photodegradation by direct photolysis, the substrate (B) absorbs light, and subsequently
through the electronically excited states singlet (1 B*) or triplet (3 B*), generates the products P1 ;
P3 or Pox .

4.2.3 Biocides selected for the study


Biocides involve a very wide range of compounds, of natural or synthetic origin. Generally, they
are classified by the target population they are supposed to kill or eradicate, as insecticides,
acaricides, fungicides, molluscicides, nematocides, rodenticides, and herbicides (Mahmoud and
Loutfy, 2012). Taking into consideration their chemical structure, one important group of pesti-
cides is constituted by the phenol derivatives. Among them, Bromoxynil (BXN) and Dichlorophen
(DCP) can be mentioned (Fig. 4.3). BXN and DCP are colorless compounds, they exhibit charac-
teristics that make them different from other phenolic pesticides: (a) BXN shows an extremely low
pKa value for the ionization of its OH group (pKa = 3.86) (Nonell et al., 1995), as compared with
phenol (pKa = 9.89) (Weast et al., 1981) and most phenolic pesticides, usually in the rage 9–11
62 J.P. Escalada et al.

Figure 4.3. Chemical structures of the pesticides: (a) Bromoxynil (BXN); (b) Dichlorophen (DCP); and (c)
two components of Abamectin (ABA): Avermectin B1a (≥80%) and Avermectin B1b (≤20%).

(Martiré and González, 2000); and (b) DCP has two phenol groups, with pKa values of 7.66 and
11.60 (Tomlin, 1994), whereas most phenolic pesticides are monohydroxylated compounds. The
presence of phenol groups and the differential ionization rates play a crucial role on kinetic and
mechanistic aspects of photodegradation process (Scully and Hoigné, 1987); these characteristics
make BXN and DCP interesting candidates for a photochemical study.
In the last decades, natural substances have been introduced as biocides, they are called biopesti-
cides. This group includes a family of macrocyclic lactones, the avermectins. They were first found
in the fermentation broth of the soil actinomycete Streptomyces avermitilis, in which eight natu-
ral components were discovered. Among the avermectins, the commercial acaricide Abamectin
(ABA), also called Avermectin B1 , is a mixture, approximately 4:1, of two colorless homologues
with the same structure (Fig. 4.3): Avermectin B1a and Avermectin B1b (Tomlin, 1994). In spite
of its extremely low acceptable daily intake (ADI) value and its observed teratogenic effects in
mouse reproduction (WHO, 1990) ABA is widely employed in veterinary medicine (Kolar et al.,
2006) and agriculture (Kamel et al., 2006), which determines the undoubted need of researches
that show the mechanism of their photodegradation.

4.2.3.1 State of the art


Phenolic pesticides and, in general hydroxyaromatic pesticides, are usually colorless, and their
phototransformation under natural or artificial conditions has received considerable attention in
the last decades (Adams et al., 2004; García, 1994; Lan et al., 2010; Tratniek and Hoigné, 1991).
In the presence of Rf, it has been reported that the parent compound phenol and several substituted
phenols can be efficiently degraded under visible irradiation, and that the photodegradation occurs
mainly through radical-driven reactions, whereas in some cases O2 (1 g )-mediated oxidations are
also involved (Görner, 2007; Gutiérrez et al., 2001; Haggi et al., 2004).
The main purpose of most of the works devoted to the environmental fading of BXN and
DCP was the identification of final products. DCP is degraded by direct photolysis yielding
4-chloro-4-hydroxymethylenediphenol in acidic anaerobic medium, and a benzoquinone deriva-
tive in aerated solutions (Mansfield and Richard, 1996). The photodegradation of DCP has also
been investigated in the form of a sand dispersion (Zertal et al., 2005), being the primary
photoproducts similar to those found in aqueous solution. The laccase-mediated oxidation of
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 63

several halogenated pesticides, including BXN and DCP, has been recently investigated (Torres
et al., 2009), looking for the reduction of their environmental impact.
A number of reports have been published about ABA persistence in nature. Thus, some degra-
dation studies using a non-discriminated combination of bio- and photo-processes indicate a
persistence of 2–4 weeks (Kamel et al., 2006; Kolar et al., 2003) either dispersed in the soil or
deposited in the field by sheep feces. In the aqueous photodegradation under natural sunlight
irradiation of Emamectin benzoate, a semisynthetic derivative of the second generation of Aver-
mectin pesticides with the structure of [4 -(epi-methylamino)-4 -deoxy]-ABA, half-life in the
range 1–22 days have been found (Musthaq et al., 1998). On the other hand, the thin film UV
photodegradation of Avermectin B1a , the main component of ABA, and of its derivative with the
former 4 -substituent, has been studied in an effort directed to know the structures of the formed
products under conditions that mimic natural layers of these pesticides on leaves (Crouch et al.,
1991; Feely et al., 1992). From these results, it was concluded that, in general, oxidation products,
geometric isomers and dealkylated compounds were formed.

4.3 METHODS FOR PHOTODEGRADATION STUDIES

In general, the experimental methods employed to quantitatively monitor photo-degradation pro-


cesses are based on absorption spectroscopy, fluorescence or HPLC techniques. Another way to
follow photodegradation is measuring oxygen consumption along the irradiation progress; this
can be done with a specific oxygen electrode (i.e. Orion 97-08). Here, the specific techniques
employed in the study of natural and artificial photodegradation of biocides will be detailed.
The overall quenching rate constant of O2 (1 g ), kt , can be determined using time resolved
phosphorescence detection (TRPD) and methodology already described by Nonell et al. (1995),
with RB as a dye sensitizer. In a homemade photolyzer with the filtered light (>350 nm, cut-off
filter) from a 150 W quartz-halogen lamp it is possible to obtain continuous aerobic photolysis
of aqueous solutions of a problem compound like biocides (B), with RB or Rf as sensitizers. The
rate constant kr for the reaction of O2 (1 g ) may be determined by oxygen consumption upon
photosensitized irradiation, according to the method described by Tratniek and Hoigné (1991)
for which the knowledge of the reactive rate constant krR for the photooxidation of a reference
compound R is required, using equation (1):
 
slope slopeR = kr [B] krR [R] (4.1)

where slope and slopeR are the respective slopes of the first-order plots of B and R’s consumption.
Generally, furfuryl alcohol (FFA) (krR = 1.2 × 108 M−1 s−1 ) is employed as a reference, which is
independent of pH.
To study the mechanism of photooxidative processes, some compounds that are specific ROS
quenchers have been employed in order to confirm/discard the involvement of different species in
a given oxidative event (Escalada et al., 2006; Silva et al., 1999). For this purpose, the quenchers
mostly used are sodium azide (NaN3 ), catalase from bovine liver (CAT) and superoxide dismutase
from bovine erythrocytes (SOD). NaN3 has been repeatedly used to study the involvement of
O2 (1 g ) because it is a well-known physical quencher of this species, with a reported rate constant,
kq , of 4.5 × 108 M−1 s−1 in water at pH 7 and 2.3 × 108 M−1 s−1 in methanol (MeOH) (Wilkinson
et al., 1995). Meanwhile, CAT reacts with H2 O2 , thereby inhibiting the photodegradation via
process (18), and the enzyme SOD disproportionates the species O•− 2 quenching process (17).
In some cases, like the time-resolved determinations of O2 (1 g ), it was necessary to extend
the lifetime of the species, this was done by using as solvents deuterated water (D2 O) or
monodeuterated methanol (MeOD) (Wilkinson et al., 1995).
In the experiments here described, direct photolysis’ quantum yields were determined
employing the actinometric method developed by Rahn (1997).
64 J.P. Escalada et al.

Figure 4.4. Spectral evolution of ABA 0.03 mM plus Rf 0.04 mM in methanol vs. Rf 0.04 mM in the same
solvent upon photoirradiation with visible light (>400 nm). The numbers in the figure indicate
the irradiation time in seconds.

4.3.1 Sensitized photoirradiation


As it was stated above, photodegradation of sunlight transparent compounds can be carried out
through photosensitized processes. Since ABA, BXN and DCP only absorb light with wave-
lengths lower than the so-called UV-B range of solar irradiation, it is clear that natural light does
not produce any chemical transformations in the molecules of these compounds. Namely, the
action of Rf as sensitizer for the natural photooxidative degradation of these biocides has been
studied.
Aerobic photosensitization produces changes on the studied biocides which clearly show the
occurrence of chemical transformations in their structures. Stationary aerobic photolysis experi-
ments were performed for ABA, BXN and DCP. Photolysis of methanol solutions containing ABA
0.03 mM plus Rf 0.04 mM or ABA (0.05 mM) plus RB (A530 = 0.5) were carried out in a Photon
Technology International (PTI) unit provided with a high pass monochromator and a 150 W Xe
lamp, irradiating with 440 ± 10 nm, or in a home-made photolyzer for non-monochromatic irra-
diation provided with a 150 W quartz-halogen lamp. In the last case, a cut-off filter at 400 nm
ensured that the light was absorbed only by the sensitizer and the spectrum was corrected by
subtracting the one of non-irradiated Rf in the same solvent (Fig. 4.4). The decomposition of
the sensitizer can be followed by the changes in the negative region of the visible spectral range
and the absorbance increase in the region of 245 nm, where also ABA absorbs. So, absorbance
changes can be used for the direct evaluation of ABA disappearance.
The photoirradiation of the mixture Rf (0.05 mM) plus a BXN or DCP (0.2–0.4 mM) in pH 7
water (>350 nm, cut off filter) produces spectral changes that can be attributed to transforma-
tions in both components of the mixtures (see the representative Fig. 4.5 for the case of BXN).
In parallel, the photoirradiations give rise to oxygen consumption. The rate of oxygen uptake at
pH 7 of the system Rf (0.04 mM) plus BXN (0.5 mM) is approximately 1.5 times faster than the
corresponding one of the system Rf (0.04 mM) plus DCP (0.5 mM) under identical experimental
conditions (Table 4.1).
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 65

Figure 4.5. Spectral evolution of a pH 7 aqueous solution of Rf (0.023 mM) plus BXN (0.47 mM), taken
vs. Rf (0.023 mM) in the same solvent, upon photoirradiation (>400 nm) under air-saturated
conditions. The numbers in the figure represent irradiation times in seconds.

Table 4.1. Rate constant values for the overall (kt ) and reactive (kr ) quenching of O2 (1 g ), relative rates of
oxygen consumption (O2 ) upon RB or Rf sensitization and quenching of 1 Rf* (1 kq ) and 3 Rf*
(3 kq ) by Bromoxynil (BXN) and Dichlorophen (DCP) in air-equilibrated solutions. Deuterated
solvents, D2 O and MeOD, were only employed in the determinations of kt values. Reported data
for phenol are included for comparison.

Compound Solvent kt (×108 ) kr (×108 ) kr /kt O2 (RB) O2 (Rf) 1 kq (×109 ) 3 kq (×109 )
(relative) (relative)

BXN H2 O pH 7 or D2 O pD 7 0.24 ± 0.04 0.16 ± 0.005 0.66 0.18 1.00


D2 O pD 12 0.25 ± 0.03
MeOH or MeOD 0.04 ± 0.002 0.02 ± 0.002 0.50 2.6 ± 0.1 2.0 ± 0.08
DCP H2 O pH 7 or D2 O pD 7 2.70 ± 0.05 0.87 ± 0.04 0.32 1.00 0.66
D2 O pD 12 7.00 ± 0.08
MeOH or MeOD 0.08 ± 0.001 0.02 ± 0.002 0.25 4.7 ± 0.1 2.1 ± 0.10
phenol H2 O pH 8 0.011 <0.011
MeOH 4.12 0.52

1 Data from Scurlock et al. (1990). 2 Data from Haggi et al. (2004).

These results strongly suggest that either Rf electronically excited states or ROS produced
through these states, or even both species operating simultaneously, are responsible for the pho-
todegradation of each pesticide. It is well known that Rf is highly reactive when its water solutions
are irradiated with visible light, due to the generation of ROS (Escalada et al., 2006; Görner, 2007;
Silva et al., 2005). In the present case, the participation of ROS was evaluated through oxygen
consumption experiments in the presence of specific ROS quenchers (Fig. 4.6). Thus, the pres-
ence of NaN3 (10 mM) or CAT (1 µg mL−1 ) neatly decreased the rate of oxygen consumption,
whereas the presence of SOD (1 µg mL−1 ) increased this rate.
At room temperature, in-situ generated or added H2 O2 (in this case 0.05 mM) to aqueous buffer
pH 7 solutions of each biocide (0.12 mM) participates in the oxidation of each pesticide, this was
confirmed by the results with CAT – Figure 4.1(18). The reaction between H2 O2 and BXN or DCP
-process (12)- was detected by the absorption or fluorescence changes, respectively. The viability
of this reaction was independently checked in the absence of light. For DCP, the spectrum was
66 J.P. Escalada et al.

Figure 4.6. Oxygen consumption for the photolysis Rf-sensitized of BXN solution (0.5 mM) pH 7 (). The
same in the presence of the inhibitors CAT (1 µg mL−1 ) (•) and SOD (1 µg mL−1 ) ().

Figure 4.7. Spectra of (a) fluorescence of DCP and (b) absorption of BXN. Both measured in absence and
presence of H2 O2 . Concentration of each reagent: ∼ 0.1 mM.

modified by the oxidation, but then it did not change for at least the next five minutes. A similar
change was observed in the absorption spectrum of BXN, a non-fluorescent compound at pH 7
due to the total ionization of its OH group (Fig. 4.7).
On the other hand, in presence of SOD the oxygen uptake increased due to a higher oxidation
rate, as a consequence of the consumption of the species O•−2 (Fig. 4.6). These species promotes
the subsequent formation of H2 O2 –process (15)–, which is concomitantly consumed in pathway
10. So, H2 O2 was involved in the process, as formerly observed with other hydroxyaromatic
compounds (Afanas’ev, 1989).
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 67

4.3.1.1 Quenching of 1 Rf* and 3 Rf*


A classical Stern-Volmer treatment allows to calculate the quenching rate constant 1 kq of the
interaction of the singlet excited state of Rf (1 Rf*) with each biocide. In order to increase the
solubility of the biocides, the quenching experiments were performed in methanol solutions of
the substrates, especially in the case of 1 Rf*, where a relatively high quencher concentration was
required. In this case, fluorescence lifetimes were evaluated with a time-correlated single pho-
ton counting technique (SPC) on an Edinburgh FL-9000 CD instrument provided with a nF900
nanosecond flash-lamp. In air-equilibrated methanol solution, Rf showed a fluorescence emis-
sion band centered at 520 nm. In the presence of biocides solutions of concentrations equal or
higher than 2 mM, the quenching of the fluorescence from 1 Rf* was detectable as a decrease
in the stationary emission intensity, but the shape of the emission spectrum did not change.
The interaction 1 Rf*-B was quantified through time-resolved methods by monitoring the fluores-
cence lifetime of 1 Rf* with different biocide concentrations. Through the classical Stern-Volmer
treatment (Eq. 4.2):

1
τ0 1 τ = 1 + 1 kq · 1 τ0 [B] (4.2)
where
1
τo : lifetime of 1 Rf* in absence of biocide;
1
τ : experimentally determined lifetime of 1 Rf* in presence of biocide;
1
kq : rate constant process (2) in Fig. 4.1) determined graphically (Fig. 4.8, Table 4.1).
The fluorescence decay of 1 Rf* in the absence and in the presence of pesticide was mono-
exponential, with 1 τo = 5.4 ns, in agreement with previously reported data (Bertolotti et al.,
1999).
The triplet excited state of Rf (3 Rf*), was generated by a 355 nm laser pulse of a nanosecond
Nd:YAG laser system (Spectron), with a 150 W Xenon lamp as analyzing light. The detection
system comprised a PTI monochromator and a red-extended photomultiplier (Hamamatsu R666).
The signal was acquired and averaged by a digital osciloscope (Hewlett-Packard 54504a). Haggi
et al. (2004) have already described the methodology of laser flash photolysis for the generation
and detection of transient species and also the apparatus required to perform it.
The decay of 3 Rf* was measured at much lower Rf concentration, typically nitrogen-saturated
aqueous 0.02 mM, and at low enough laser energy in order to avoid self-quenching and triplet-
triplet annihilation, respectively. The disappearance of 3 Rf* was monitored by the first-order decay
of the absorbance at 670 nm, a wavelength where the interference from other possible species
is negligible. The lifetime of 3 Rf* appreciably decreased in the presence of these biocides –
Figure 4.1(4). As before, a Stern-Volmer treatment of the triplet quenching was applied (Fig. 4.9),
using the expression of Equation (4.3):

3
τ0 3 τ = 1 + 3 kq · 3 τ0 [B] (4.3)
where
3
τo : lifetime of 3 Rf* in absence of biocide;
3
τ : experimentally determined lifetime of 3 Rf* in presence of biocide;
3
kq : bimolecular rate constant.
The quenching of 3 Rf* by electron-donating compounds such as phenols and polyhydroxy-
benzenes occurs through an electron transfer process – Figure 4.1(9)–, with high values of the
respective rate constant 3 kq , close to 109 M−1 s−1 (Haggi et al., 2004). The thermodynamic fea-
sibility of process (9) can be evaluated by means of the Gibbs free energy for electron transfer
(Eq. 4.4):

ET G0 = E0(B/B+ ) − E0(Rf /Rf − ) − ERf ∗ + C (4.4)


where
E0(Rf /Rf − ) : standard electrode potential of the acceptor Rf (−0.80 V),
ERf ∗ : 3 Rf* energy (2.17 eV),
68 J.P. Escalada et al.

Figure 4.8. Stern-Volmer plot for the quenching of 1 Rf* in MeOH by BXN and DCP. τo and τ refer to 1 Rf*
lifetimes in absence and in presence of pesticide, respectively.

Figure 4.9. Stern-Volmer plots for the quenching of 3 Rf* by BXN in MeOH. τo and τ refer to 3 Rf* lifetimes
in the absence and in the presence of BXN, respectively.

C: Coulombic energy term (−0.06 V) (Porcal et al., 2003) and


E0(B/B+ ) : standard electrode potential of each pesticide.
Under aerobic conditions, a cascade of processes can occur after process (9) (Fig. 4.1),
some of them generating ROS. Thus, Rf•− can undergo a fast protonation to RfH•−
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 69

Figure 4.10. Transient absorption spectra of Rf (0.05 mM) in argon-saturated MeOH without pesticide
(1 µs after the laser pulse) and plus DCP or BXN (both 5 µs after the laser pulse).

•−
Figure 4.1(14)- (Lu et al., 2004), or can transfer an electron to O2 (3 − g ) yielding O2 and

regenerating ground state Rf, process (13). Then, the species RfH can undergo a bimolecu-
lar disprotonation reaction yielding fully reduced Rf (RfH2 ) and Rf -process (15). In the presence
•+ •–
of O2 (3 −
g ), RfH2 can be reoxidized, giving rise to RfH2 and O2 and, finally, to Rf and H2 O2 –
•–
process (16). Both oxidative species O2 and H2 O2 may react with the pesticides in the medium.
The generation of both species, Rf•− and RfH• occurs from several hydroxyaromatic substrates
of environmental and biological importance (Haggi et al., 2002; Pajares et al., 2010).
The species B•+ , formed in process (9), either produces the decay or regeneration of the
pesticide through ulterior radical recombination reactions. The alternative generation of O•− 2 by
direct interaction of 3 Rf* with O2 (3 −
g ) has not been taken into account in the present discussion
because of the very low quantum yield of process (10) (Krishna et al., 1991).
The determined values of 1 kq and 3 kq for BXN or DCP (Table 4.1) are close to the diffusion
limit and similar to those reported for phenol (Haggi et al., 2004).
In the absence of a biocide, a spectrum similar to the expected one for 3 Rf* was observed after
the laser pulse (Fig. 4.10) (Gutiérrez et al., 2001; Haggi et al., 2002; Pajares et al., 2001), while
in the presence of a biocide (0.6 mM, approximately 95% 3 Rf* quenched by the pesticide) the
shape of the long-lived absorption is in good agreement with that reported for the semiquinone
radical RfH• , formed from the radical anion Rf•− – Figure 4.1(14) – (Barbieri et al., 2008; Lu
et al., 2004). The experimental E0(B/B+) values obtained were 0.71 V for BXN and 0.62 V for
DCP (Bard and Faulkner, 2001). The so-calculated ET G0 values, lower than −0.71 eV for both
pesticides, indicate that process (9) may be operative and, consequently, that the species RfH•
could be spontaneously formed – Figure 4.1(14).
The prevalence of either the electron transfer process from a biocide or the energy transfer
process from 3 Rf* to O2 (3 −
g ) (process (9) vs. process (5), respectively) can be discussed in kinetic
terms. Considering the described kET value of process (5) in H2 O (7 × 108 M−1 s−1 , i.e. 1/9 of
the diffusional value) (Koizumi et al., 1978) and a 3 kq value for both biocides of 2 × 109 M−1 s−1
70 J.P. Escalada et al.

Figure 4.11. Photoirradiation processes of Rf alone and in the presence of ABA 0.03 mM, follow by the
variations in absorbance at 245 nm. Inset: Stern-Volmer plot for the stationary quenching of
the fluorescence of riboflavin 0.06 mM by ABA, in MeOH.

(Table 4.1), it can be deduced that, for the same concentrations of B and dissolved O2 (3 − g ), the
rate for the radical-mediated pathway, process (9), is approximately three times higher than the
O2 (1 g )-generation, process (5). In practical terms, this indicates the operation of both processes,
(9) and (5) (Fig. 4.1).
There is a narrow but quite reproducible disparity between the rates of spectral changes at
245 nm during the irradiation of a solution of Rf 0.04 mM in methanol, in the absence and in
the presence of ABA 0.03 mM (Fig. 4.11). Taking into account that only the sensitizer absorbs
the exciting light, the decrease in the rate of Rf decomposition in the presence of ABA could be
due to: (i) quenching of Rf excited states by ABA (the photodegradation of Rf under visible light
irradiation proceeds mainly through the triplet state, 3 Rf∗ ) (Heelis, 1982), and/or (ii) decrease in
the absorbance of the spectral component of ABA at 245 nm, due to photoreaction, producing a
decrease in the observed rate of overall (Rf + ABA) absorbance increase at this wavelength.
Rf presents an intense fluorescence emission band centered at 515 nm, with a reported fluo-
rescence quantum yield of 0.25 (Heelis, 1982). The stationary fluorescence of air equilibrated
aqueous solutions of Rf at 25 ± 1◦ C, excited at a wavelength of 445 nm and emission at 515 nm
was registered in a Rf 5301-PC Shimadzu spectrofluorimeter. Figure 4.11, inset, shows the Stern-
Volmer plot for the quenching of the stationary emission of Rf in methanol. As it can be seen, the
quenching by ABA up to 0.05 mM, a concentration higher than that employed in the photolysis
experiments, is negligible. In parallel experiments, the addition of ABA (up to 0.1 mM) to a N2 -
saturated solution of Rf 0.01 mM in methanol did not affect the 3 Rf* lifetime of 15 µs observed
in the absence of ABA, a value in agreement with literature data (Heelis, 1982). Besides, the
transient absorption spectrum of the solution taken at 0.1 µs after the laser pulse, attributed to the
neutral 3 Rf* species (Haggi et al., 2004), does not change in the presence of ABA.
In synthesis, the results on the potential quenching of Rf singlet and triplet excited states by
ABA, obtained by means of stationary fluorescence and laser flash photolysis, indicate absence
of interaction between ABA and the excited states of Rf, at least within the tested concentration
range of ABA, 0.05–0.10 mM, values higher than those employed in the experiments shown in
Figure 4.4.
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 71

Figure 4.12. (a) UV-Vis absorption spectrum of MeOH solutions of RB (A560 = 0.46) plusABA (0.025 mM)
taken vs. RB, during the irradiation; the numbers on the spectra represent irradiation times in
seconds. (b) Oxygen uptake of a solution of RB (A560 = 0.46) plus ABA 1 mM in MeOH with
and without NaN3 0.1 mM.

4.3.1.2 Quenching of O2 (1 g )
The possible participation of O2 (1 g ) in the photodegradation of the biocides can be evalu-
ated employing RB, a well-known and almost exclusive O2 (1 g ) generator, which in methanol
presents a quantum yield of 0.81 (Amat-Guerri et al., 1990) and in water of 0.76 (Neckers,
1989; Wilkinson et al., 1995). RB was chosen as a sensitizer in order to focalize on the potential
reaction of each pesticides with O2 (1 g ), avoiding possible interferences due to interactions of
the substrates with Rf electronically excited states. The spectral changes observed upon aerobic
photosensitization clearly show the occurrence of chemical transformations in biocides. Upon
visible-light photoirradiation of methanol solution of ABA plus RB (A560 = 0.46) as photosen-
sitizer, the photodegradation can be followed by the changes in ABA spectrum and the rate of
oxygen consumption. In the presence of NaN3 0.1 mM a lowering of the rate of oxygen consump-
tion was detected, confirming that the photodegradation process depends on O2 (1 g ) generation
(Fig. 4.12).
When analogous experiments were performed in aqueous media, similar processes took place;
Figure 4.13 shows the small changes produced in the biocides spectra when their aqueous pH 7
solutions were irradiated with visible light in the presence of RB; also oxygen consumption was
observed in water solutions.
The rate of oxygen uptake in the system RB (A530 = 0.5) plus DCP (0.5 mM) was approximately
5.5 times faster than in the system RB (A530 = 0.5) plus BXN (0.5 mM), measured under identical
experimental conditions (Table 4.1). As a comparison, the system RB (A530 = 0.5) plus phenol
(0.5 mM) practically did not consume oxygen at all, within the same temporal window employed
for the biocides. These results strongly suggest some degree of interaction of O2 (1 g ) with the
biocide. The decay kinetics of O2 (1 g ) phosphorescence was first order, and the lifetime agreed
well with literature data (Wilkinson et al., 1995).
The addition of a biocide as a quencher leads to a decrease of the O2 (1 g ) lifetime, unam-
biguously confirming the interaction of the pesticides with this oxidative species. The kt values
72 J.P. Escalada et al.

Figure 4.13. UV-Vis absorption spectrum of aqueous solutions of RB (A560 = 0.46) plus ABA (0.025 mM)
taken vs. RB (A560 = 0.46), during the irradiation; the numbers on the spectra represent
irradiation times in seconds.

(Table 4.1) were graphically obtained in deuterated water (D2 O) at pD 71 or pD 122 and in
deuterated MeOH (MeOD) (Fig. 4.12) through Equation (4.5):

τ0 τ = 1 + kt τ0 [B] (4.5)
1 1
where τ : O2 ( g ) lifetimes in the presence of a biocide and τ0 : O2 ( g ) lifetimes in the
absence of a biocide.
Highly alkaline medium and methanol were employed as solvents in order to evaluate possible
effects of pH and solvent polarity on the rate constants of the interaction O2 (1 g )-B (Table 4.1).
Both kt and kr values for both biocides are significantly higher than those described for phenol
(Scully and Hoigné, 1987). Furthermore, according to the data, the interaction O2 (1 g )-phenol
appears as almost totally physical in nature (Table 4.1). Scurlock et al. (1990) have also reported
the absence of observable effective O2 (1 g )-mediated phenol oxidation in H2 O.
The rate constant kr – Figure 4.1(8) – was determined by the above mentioned acti-
nometric method, monitoring oxygen photoconsumption (Fig. 4.15). The O2 (1 g )-mediated
photooxidation quantum efficiency
r , is express by Equation (4.6),


r = kr [B] (kd + kr [B]) (4.6)
particularly in natural environments. Quantum efficiency is difficult to determine in natural
environments due to the fact that the actual concentration of the photooxidizable substrates,
represented by B in this case (García, 1994), needs to be known beforehand.
A simpler and useful approach is the evaluation of the kr /kt ratio (Table 4.1), which indicates
the fraction of overall quenching of O2 (1 g ) by the substrate that effectively leads to a chemical
transformation.

1 pD = 7: Measured at molar deuterium concentration [D] = 1 × 10−7


2 pD = 12: Measured at molar deuterium concentration [D] = 1 × 10−12
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 73

Figure 4.14. Stern-Volmer plot for the quenching of O2 (1 g ) phosphorescence by DCP in D2 O, pD 12


(left and bottom axes), and in MeOD solution (right and top axes).

Rf is an excellent model sensitizer for the photooxidation of different contaminants in natural


environments, since under visible light irradiation it produces O2 (1 g ) and O− 2 − processes (5)
and (10) (Fig. 4.1). With Rf-sensitization, the different trend of the relative rates of oxygen uptake
(Table 4.1), as compared to those obtained under RB-sensitization, suggests the operation of a
more complex mechanism, at least different from a simple O2 (1 g )-mediated one. It is known
that Rf in methanol produces mainly O2 (1 g ), and also some O− 2 , with quantum yields of 0.48 and
0.009 respectively (Chacón et al., 1988). In principle, the participation of this radical anion in the
photodegradation of ABA – process (11) – is discarded due to the very low quantum yield for the
O− 1
2 generation. Hence, the remaining possibility is O2 ( g ) mediated reactions – Figure 4.1(5).
Oxygen consumption was observed in the visible-light photoirradiation of solutions of Rf
0.05 mM plus ABA 0.1 mM in MeOH-H2 O 3:1 v/v. The photolysis was accompanied by a fading
of the sensitizer, as directly observed by the marked bleaching of the solution, pointing again
to the chemical involvement of Rf in the photooxygenation process. In the presence of electron
donating species (eD), O− 2 can also appear through processes (9) and (13) (Fig. 4.1), with S = Rf
(Escalada et al., 2006; García and Amat-Guerri, 2005). Evidently, the two Avermetins hardly act
as electron donating agents, as suggested by the high oxidation potential that corresponds to their
structures (Weinberg and Weinberg, 1968). It is well known that Rf is involved in redox processes
in nature through different mechanisms, including its own O2 (1 g )-mediated photooxidation –
Figure 4.1(8)–, in which Rf competes with ABA for the reaction with O2 (1 g ), where for Rf
a value of k = 6 × 107 M−1 s−1 in MeOH was reported (Chacón et al., 1988). Rather low kt
values (the sum of kr + kq ), – processes 7 and 8 respectively (Fig. 4.1) – were found for the overall
quenching process of O2 (1 g ) with ABA (Fig. 4.16). In MeOD, kt = 5.5 ± 0.2 × 105 M−1 s−1 was
determined, a value similar to those generally reported for kr values of reactions of O2 (1 g ) with
isolated double bonds to form allylic hydroperoxides, the so-called Schenck ene reaction (Gollnick
and Kuhn, 1979), or with conjugated dienes yielding unsaturated endoperoxydes (Clennan, 2000),
as in the cases of 2,5-dimethyl-2,4-hexadiene, with kr = 2 × 106 M−1 s− in MeOH (Wilkinson
et al., 1995), or trans,trans-2,4-hexadiene, with kr = 1.1 × 105 M−1 s−1 in chloroform (Frimer,
1985). Hence, in the case of ABA, the probable targets for O2 (1 g ) addition, with subsequent ABA
74 J.P. Escalada et al.

Figure 4.15. Oxygen uptake in pH 7 H2 O by DCP and by FFA, in both with RB as a sensitizer (A548 = 0.5).

Figure 4.16. Stern-Volmer plot for the quenching of O2 (1 g ) phosphorescence by ABA in MeOD.

degradation, could be the three isolated double bonds at C3, C14 and C22, and the conjugated
diene group at C8-C11.
The quenching ability of biocides towards O2 (1 g ) is moderate-to-high, according to the
respective kt and kr values. For phenols and polyhydroxybenzenes (García, 1994; Miskoski et al.,
2005), the accepted mechanism of the quenching is the initial formation of an encounter complex
[O2 (1 g )-substrate] with partial charge-transfer character, from which an irreversible electron
transfer process would yield the photooxidation products. The formation of this complex depends
on the electron-donor ability of the substrate, and is favored in polar solvents. In phenols, the
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 75

Figure 4.17. Evolution during direct photoirradiation (254 nm) of the absorption spectrum of (a) ABA and
(b) BXN, both in MeOH solutions. Inserted numbers indicate seconds of irradiation.

corresponding OH-ionized forms are the most photooxidizable species. In BXN, the electron
withdrawing groups in the benzene ring decrease the pKa value, selectively favoring the photoox-
idation – Figure 4.1(14). In DCP, the presence of two phenol groups in the molecule gives rise to a
relatively high value of the rate constant kt . In spite of this, the observed kr /kt ratio in this pesticide
indicates less efficient oxidation than in the case of BXN. This result can be a consequence of the
partially-ionized and non-ionized forms of the respective OH groups in the DCP molecule at pH
7. The formation of the encounter-complex intermediate in these two compounds is supported by
the dependence of the kinetic data on the pH of the medium and on the solvent polarity.
The rate constant kt for BXN in pD 12 solution is the same as that in pD 7, due to the phenol
group total ionization, whereas in DCP a substantial kt increase is observed at higher pD value.
In DCP at pD 12, both OH groups are almost totally ionized, whereas at pH 7 only one of the
OH groups is partially ionized. In MeOD solution, both kt and kr values dramatically decrease
for BXN and DCP (Table 4.1). This dependence is interpreted as a solvent polarity effect on the
charge-transfer mediated encounter complex.

4.3.2 Direct photolysis of ABA, BXN and DCP


The quantum yield for the direct photolysis with radiation over 300 nm of BXN (Kochany et al.,
1990; Millet et al., 1998) and DCP (Mansfield and Richard, 1996) are extremely low, higher
energy is necessary in order to obtain a significant photodegradation. On the other hand, the
photoirradiation with monochromatic light of 254 nm of air-equilibrated solutions of methanol
solutions of the biocides (A254 > 2) produces spectral changes (Fig. 4.17) and presents high quan-
tum yields. This was achieved with an Osram Germicide lamp (15 W) in a magnetically-stirred
1 × 1 cm fluorescence quartz cuvette. For ABA 0.03 mM it was determined
r = 0.23 ± 0.02,
for BXN was
r = 0.18 ± 0.02, for BXN methanol solution nitrogen-saturated
r = 0.13 ± 0.02
and for anaerobic solutions of DCP in methanol
r = 0.67 ± 0.02 was obtained. All these results
indicated that under these conditions and with high energy radiation, the direct photolysis is a
very efficient method for the degradation of these three biocides, especially for DCP.
r was
graphically obtained from the respective rates of biocides and actinometer photolysis (Fig. 4.18)
76 J.P. Escalada et al.

Figure 4.18. Number of reacting molecules as a function of irradiation time with 254 nm light, for acti-
nometer, ABA, BXN (left and bottom axes) and DCP (top and right axes), employed for the
determination of quantum yield of direct photodegradation of the pesticides.

employing the iodate-iodide actinometer (Rahn, 1997). The photolysis energy, approximately
470 kJ Einstein−1 , is strong enough to break C-H and C-C bonds.
The rates of ABA photoconsumption were practically the same, within experimental error,
when the photolysis was carried out in N2 -equilibrated conditions. Also for BXN, the slight
difference between air and nitrogen saturated solutions is not sufficient to dismiss that a parallel
oxygen mediated oxidation process is also taking place.
It must be pointed out that, by this method, it was not possible to measure the quantum yield
of DCP in aerobic condition, due to the occurrence of photoproducts that absorb in the same
spectral region than the pesticide. The products may be quinine compounds, very frequent in the
oxidation of phenol derivatives (Wilkinson et al., 1993), already reported by Zertal et al. (2005)
for DCP.

4.4 CONCLUSIONS

In environmental sciences, the interest on the application of the photochemical degradation of


pesticides has grown enormously in the last decades. As explained above, the phenolic deriva-
tives ABA, BXN and DCP can be decomposed either by direct or by indirect photolysis. In
order to get more insight in these processes, some questions should be answered. Is the natural
degradation better than the artificial? Or, is it the other way round? Do both, direct and indirect
photodegradation, have the same efficiency? Can both processes take place in every system or
media?
The authors think that the experiments presented in the previous pages help to shed some light
over these inquiries.
Generally, the degradation efficiency of organic pollutants, among them ABA, BXN and DCP,
can be improved by using indirect photochemical processes. One common indirect photolysis
mechanism is by singlet oxygen generation, which consumes oxygen through electrophilic attack
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 77

on the organic components. The Rf-photosensitized oxidation of biocides occurs through a mech-
anism mediated by H2 O2 and O2 (1 g ). Through the O2 (1 g )-mediated process, these pesticides
are more easily degraded than phenol—a compound hardly degradable by photooxidation—, due
to much lower pKa values of the hydroxyl groups in these compounds. This opens the possibil-
ity for designing photodegradation-time-tunable phenolic pesticides by an adequate substitution
pattern in the phenol molecule. The new substituent has to achieve for the OH group a pKa value
low enough to improve the natural degradation.
Biocides undergo photodegradation with relatively high quantum yield when irradiated with
254-nm light. The energetic input rounds 470 KJ, a value strong enough as to break down
C-C single bonds. Dissolved oxygen concentration does not affect the reaction rate, so oxygen-
mediated oxidative process can be discarded. Nevertheless, in a contaminated medium and in the
presence of other efficient to moderated competitors for the reaction with O2 (1 g ), ABA possess
limited possibilities of photosensitized degradation induced by environmental light.
Indirect photolysis in lagoons under solar radiation has been proposed for the treatment of
pesticide-laden agricultural runoff, pesticide container rinses, and industrial wastes. It must
be taken into account that, photolysis lagoons will require optimum oxygen input to achieve
economical treatment.
Direct photodecomposition processes can be used whenever organic pesticides absorb light
sufficiently in the UV region (250–300 nm) (Coly and Aaron, 1994; Patria et al., 1995). However,
several factors often limit the efficiency of these photoprocesses in aquatic environments: the
transparency of natural waters, pesticide molar extinction coefficient and water solubility, solar
light absorption and pH value of water, which may shift the pesticide absorption maximum and
produce significant effects on photohydrolysis kinetics. For example, in the case of the direct
photolysis of aromatic pesticides, the conversion yield has been shown to range from 18 to 99.5%
for reaction times of 40–70 min, according to the molecular structure of compounds and the nature
of the solvent (Coly and Aaron, 1994; 1998).
The direct photoirradiation has high quantum yield when irradiated with ultraviolet light. This
kind of energetically-expensive treatment for the biocides degradation could be of interest in case
of deposits of pesticides for which a rapid degradation is needed.
It can be concluded that the main advantage of direct photolysis is the great efficiency of the
process. Among the disadvantages it can be said the requirement of a huge amount of energy,
leading to a process that may be more expensive. On the other hand, the sensitized photodegra-
dation uses as energy supply the solar light, which is obtained for free. It must be said that,
although artificial photodegradation is cheaper than natural, the possibility of employment of Rf
as a dye sensitizer for the programmed removal of pollutants is not the best choice because it is
a less efficient process due to the fact that the reactions involved are too slow. For that reason,
it is preferable to employ RB, which is an efficient artificial ROS generator. Another point that
must be taken into account is that sensitizers are degraded as well, thus making the control of
the variables in the indirect photoprocess more difficult than in the direct one. Nevertheless, as
Rf is unavoidably present in practically all natural water courses and lakes, the slowness of the
degradative reactions could be compensated with time of environmental light irradiation.

REFERENCES

Adams, W.A., Bakker, M.G., Macías, T. & Jefcoat, I.A.: Synthesis and characterization of mesoporous silica
films encapsulating titanium dioxide particles. Photodegradation of 2,4-dichlorophenol. J. Hazard. Mater.
B 112, (2004), pp. 253–259.
Afanas’ev, I.B.: Superoxide Ion: Chemistry and Biological Implications, vol. 1, CRC Press, Boca Raton, FL,
1989.
Amat-Guerri, F., López-González, M.M.C., Martínez-Utrilla, R. & Sastre, R.: Singlet oxygen photogenera-
tion by ionized and un-ionized derivatives of Rose Bengal and Eosin Y in diluted solutions. J. Photochem.
Photobiol. A: Chem. 53 (1990), pp. 199–210.
78 J.P. Escalada et al.

Barbieri,Y., Massad, W.A., Díaz, D., Sanz, J., Amat-Guerri, F. & García, N.A.: Photodegradation of Bisphenol
A and related compounds under natural-like conditions in the presence of riboflavin. Kinetics, mechanism
and photoproducts. Chemosphere 73 (2008), pp. 564–571.
Bard, A.J. & Faulkner, L.R.: Electrochemical methods: Fundamentals and applications. Wiley, New York,
2001.
Benassi, C.A., Scoffone, E., Galiazzo, G. & Jori, G.: Proflavin-sensitized photooxidation of tryptophan and
related peptides. Photochem. Photobiol. 28 (1967), pp. 857–862.
Bertolotti, S.G., Previtali, C.M., Rufs, A.M. & Encinas, M.V.: Riboflavin/triethanolamine as photoinitiator
system of vinyl polymerization. A mechanistic study by laser flash photolysis. Macromolecules 32 (1999),
pp. 2920–2924.
Chacón, J. N., McLearie, J. & Sinclair, R.S.: Singlet oxygen yields and radical contributions in the dye-
sensitized photo-oxidation in methanol of esters of polyunsatured fatty acids (oleic, linoleic, linolenic
and arachidonic). Photochem. Photobiol. 47 (1988), pp. 647–656.
Coly, A. & Aaron, J.J.: Fluorometric-determination of aromatic pesticides in technical formulations - effects
of solvent and of ultraviolet photolysis. Talanta 41 (1994), pp. 1475–1480.
Coly, A. & Aaron, J.J.: Cyclodextrin-enhanced fluorescence and photochemically-induced fluorescence
determination of 5 aromatic pesticides in water. Anal. Chim. Acta 360 (1998), pp. 129–141.
Crouch, L.S., Feely, W.F., Arison, B.H., VandenHeuvel, W.J.A., Colwell, L.F., Stearns, R.A., Kline, W.F. &
Wislocki, P.G.: Photodegradation of Avermectin B1a thin films on glass. J. Agric. Food Chem. 39 (1991),
pp. 1310–1319.
De la Rochette, A., Silva, E., Bilourez-Aragon, I., Mancini, A., Edwards, A.-M. & Molière, P.: Riboflavin
photodegradation and photosensitizing effect is highly dependent on oxygen and ascorbate concentrations.
Photochem. Photobiol. 72 (2000), pp. 815–820.
Escalada, J.P., Pajares, A, Gianotti, J., Massad, W.A., Bertolotti, S., Amat-Guerri, F. & García, N.A.: Dye-
sensitized photodegradation of the fungicide carbendazim and related benzimidazoles. Chemosphere 65
(2006), pp. 237–244.
Escalada, J.P., Gianotti, J., Pajares, A, Massad, W.A., Bertolotti, S., Amat-Guerri, F. & García, N.A.:
Photodegradation of the acaricide abamectin. A kinetic study. J. Agr. Food. Chem. 56 (2008), pp. 7355–
7359.
Escalada, J.P., Pajares, A, Gianotti, J., Biasutti, A., Criado, S., Molina, P. & Massad, W.A., Amat-Guerri,
F. & García, N.A.: Photosensitized degradation in water of the phenolic pesticides Bromoxynil and
Dichlorophen in the presence of Riboflavin, as a model of their natural photodecomposition in the
environment. J. Hazard. Mat. 186 (2011), pp. 466–472.
Feely, W.F., Crouch, L.S., Arison, B.H., VandenHeuvel, W.J.A., Colwell, L.F. & Wislocki, P. G.: Photodegra-
dation of 4 -(epimethylamino)-4 -deoxyavermectin B1a thin films on glass. J. Agric. Food Chem. 40
(1992), pp. 691–696.
Gao, J., Wang, Y., Gao, B., Wu, L. & Chen, H.: Environmental fate and transport of pesticides. In: H.S.
Rathore & L.M.L. Nollet (eds): Pesticides: Evaluation of environmental pollution. CRC Press, Taylor &
Francis Group, Boca Raton, FL, USA, 2012, pp. 29–45.
García, N.A.: Singlet molecular oxygen-mediated photodegradation of aquatic phenolic pollutants. A kinetic
and mechanistic overview J. Photochem. Photobiol. B: Biol. 22 (1994), pp. 185–196.
Görner, H.: Oxygen uptake after electron transfer form amines, amino acids and ascorbic acid to
triplet flavines in air-saturated aqueous solution. J. Photochem. Photobiol. B: Biol. 87 (2007),
pp. 73–80.
Gutiérrez, I., Criado, S., Bertolotti, S. & García, N.A.: Dark and photoinduced interactions between Trolox,
a polar-solvent-soluble model for vitamin E, and riboflavin. J. Photochem. Photobiol. B: Biol. 62 (2001),
pp. 133–139.
Haggi, E., Bertolotti, S., Miskoski, S., Amat-Guerri, F. & García, N.A.: Environmental photodegradation
of pyrimidine fungicides. Kinetics of the visible-light-promoted interactions between riboflavin and 2-
amino-4-hydroxy-6-methylpyrimidine. Can. J. Chem. 80 (2002), pp. 62–67.
Haggi, E., Bertolotti, S. & García, N.A.: Modeling the environmental degradation of water contam-
inants. Kinetics and mechanism of the riboflavin-sensitised-photooxidation of phenolic compounds.
Chemosphere 55 (2004), pp. 1501–1507.
Heelis, P.F.: The photophysical and photochemical properties of flavins (isoalloxazines). Chem. Soc. Rev. 11
(1982), pp. 15–39.
Kamel, A., Al-Dorsay, S., Ibrahim, A. & Iqbal, A.: Degradation of acaricides abamectin, flufenoxuron and
amitraz on Saudi Arabia dates. Food Chem. 100 (2006), pp. 1590–1593.
Kinetic aspects of the photodegradation of phenolic and lactonic biocides 79

Kochany, J., Choudhry, G.G., Barrie & Webster, G.R.: Photochemistry of halogenated benzene derivatives.
Part IX. Environmental aquatic phototransformation of bromoxynil (3,5-dibromo-4-hydroxybenzonitrile).
Pest. Sci. 28 (1990), pp. 69–81.
Koizumi, M., Kato, S., Mataga, N., Matsuura & T., Isui, I.: Photosensitized reactions. Kagakudogin
Publishing Co., Kyoto, Japan, 1978.
Kolar, L., Marc, I., Kužner, J., Flajs, V.C., Pogaènik M. & Kožuh, E.: Degradation of abamectin in soil from
sheep grazing pasture. Toxicol. Lett. 144 (2003), pp. 173–173.
Kolar, L., Flajs, V. C., Kužner, J., Mark, I., Pogaènic, M., Bidovec, A., van Gestel, C.A.M. & Eržen, N.K.:
Time profile of abamectin and doramectin excretion and degradation in sheep faeces. Environ. Poll. 144
(2006), pp. 197–202.
Krishna, C.M., Uppuluri, S., Riesz, P., Zigler, J.S. & Balasubramanian, D.: A study of photodynamic
efficiencies of some eye lens constitutes. Photochem. Photobiol. 54 (1991), pp. 51–56.
Lan, Q., Li, F.-B., Sun, C.-X., Liu, C.-S, & Li, X.Z.: Heterogeneous photodegradation of pentachlorophenol
and iron cycling with goethite, hematite and oxalate under UVA illumination. J. Hazard. Mat. 174 (2010),
pp. 64–70.
Lu, C., Bucher, G. & Sander, W.: Photoinduced interactions between oxidized and reduced lipoic acid and
riboflavin (Vitamin B2). Chem. Phys. Chem. 5 (2004), pp. 47–56.
Mahmoud, M.F. & Loutfy, N.: Uses and environmental pollution biocides. In: H.S. Rathore & L.M.L. Nollet
(eds): Pesticides: Evaluation of environmental pollution. CRC Press, Taylor & Francis Group, Boca
Ratón, FL, USA, 2012, pp. 3–25.
Mansfield, E. & Richard, C.: Phototransformation of dichlorophen in aqueous phase. Pest. Sci. 48 (1996),
pp. 73–76.
Mártire, D.O. & González, M.C.: Quantitative structure-activity relationship (QSAR) for reactions of singlet
oxygen with phenols. Recent Res. Devel. Photochem. Photobiol. 4 (2000), pp. 271–280.
Millet, M., Palm, W.-U. & Zetsch, C.: Abiotic degradation of halobenzonitriles: Investigations of the
photolysis in solution. Ecotoxicol. Environ. Saf. 41 (1998), pp. 44–50.
Momzikoff, A., Santus, R. & Giraud, M.: A study of the photosensitizing properties of seawater. Mar. Chem.
12 (1983), pp. 1–14.
Mushtaq, M., Chukwudebe, A.C., Wrzesinski, C., Allen, L.R.S., Luffer-Atlas, D. & Arison, B.H.: Pho-
todegradation of Emamectin benzoate in aqueous solutions. J. Agric. Food Chem. 46 (1998), pp.
1181–1191.
Neckers, D.C.: Rose Bengal Review Article. J. Photochem. Photobiol. A: Chem. 47 (1989), pp. 1–29.
Nonell, S., Moncayo L., Trull F., Amat-Guerri F., Lissi, E.A., Soltermann A.T., Criado S. & García N.A.:
Solvent influence on the kinetics of the photodynamic degradation of trolox, a water soluble model
compound for vitamin E. J. Photochem. Photobiol. B: Biol. 29 (1995), pp. 157–168.
Nubbe, M.E., Adams, V.D. & Moore,W.M.: The direct and sensitized photo-oxidation of hexachlorocyclopen-
tadiene. Water Res. 29 (1995), pp. 1287–1293.
Pajares, A., Gianotti, J., Stettler, G., Bertolotti, S., Criado, S., Posadaz, A., Amat-Guerri, F. & García
N.A.: Modeling the natural photodegradation of water contaminants. A kinetic study on the light-induced
aerobic interactions between riboflavin and 4-hydroxypyridine. J. Photochem. Photobiol. A: Chem. 139
(2001), pp. 199–204.
Pajares, A., Bregliani, M., Montaña, P., Criado, S., Massad, W., Gianotti, J., Gutiérrez, I. & García N.A.:
Visible-light promoted photoprocesses on aqueous gallic acid in the presence of riboflavin. Kinetics and
mechanism. J. Photochem. Photobiol. A: Chem. 209 (2010), pp. 89–94.
Patria, L., Merlet, N. & Dore M.: Degradation D’un Herbicide, La Flurochloridone, Par Hydrolyse et Pho-
tolyse Hydrolysis and Photodegradation of the Herbicide Flurochloridone. Environ. Technol. 16 (1995),
pp. 315–327.
Porcal, G., Bertolotti, S.G., Previtali, C.M. & Encinas, M.V.: Electron transfer quenching of singlet and
triplet excited states of flavins and lumichrome by aromatic and aliphatic electron donors. Phys. Chem. 5
(2003) pp. 4123−4128.
Rahn, R.O.: Potassium iodide as a chemical actinometer for 254 nm radiation: Use of lodate as an electron
scavenger. Photochem. Photobiol. 66 (1997), pp. 450–455.
Ross, D.R. & Crosby, D.G.: Photolysis of ethylenethiourea. J. Agr. Food Chem. 21 (1976),
pp. 335–337.
Scurlock, R., Rougee M. & Bensasson R.B.: Redox properties of phenols, their relationships to singlet
oxygen quenching and to their inhibitory effects on benzo(a)pyrene-induced neoplasia. Free Rad. Res.
Commun. 8 (1990), pp. 251–258.
80 J.P. Escalada et al.

Silber, J., Silbera, N. & Previtali, C.M.: Photoreactions of riboflavin in the presence of 2,4-
dichlorophenoxyacetic acid (2,4-D). J. Agr. Food Chem. 24 (1976), pp. 679–680.
Silva, E., Edwards, A.M. & Pacheco, D.: Visible light-induced photooxidation of glucose sensitized by
riboflavin. J. Nutr. Biochem. 10 (1999), pp. 181–185.
Silva, E., Herrera, L., Edwards, A.M., De La Fuente, J. & Lissi, E.: Enhancement of riboflavin-mediated
photo-oxidation of glucose 6-phosphate dehydrogenase by uronic acid. Photochem. Photobiol. 81 (2005),
pp. 206–211.
Tomlin, C.: The pesticide manual. British Crop Protection Council and The Royal Society of Chemistry,
London, UK, 1994.
Scully, F.E. & Hoigné, J.: Rate constants for reactions of singlet oxygen with phenols and other compounds
in water. Chemosphere 4 (1987), pp. 681–694.
Tratniek, P.G. & Hoigné, J.: Oxidation of substituted phenols in the environment: A QSAR analysis of rate
constants for reaction with singlet oxygen. Environ. Sci. Technol. 25 (1991), pp. 1956–1964.
Wilkinson, F., Helman, W. P. & Ross, A.B.: Quantum yields for the photosensitized formation of the lowest
electronically excited singlet-state of molecular oxygen in solution. J. Phy. Chem. Ref. Data 22 (1993),
pp. 113–262.
Wilkinson, F., Helman, W.P. & Ross, A.: Rate constants for the decay and reactions of the lowest electronically
excited state of molecular oxygen in solution. An extended and revised compilation. J. Phys. Chem. Ref.
Data 24 (1995), pp. 663–1021.
World Health Organization: Principles of toxicological assessment of pesticide residues in food. Environ.
Health Criterium No 104, Geneva, Switzerland, 1990.
Zertal, A., Jacquet, M., Lavédrinec, B. & Sehilia, T.: Photodegradation of chlorinated pesticides dispersed
on sand. Chemosphere 58 (2005), pp. 1431–1437.
CHAPTER 5

Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst


in a recirculating batch reactor

Natalia Inchaurrondo, Josep Font & Patricia Haure

5.1 INTRODUCTION

The freshwater available in the world continues to decrease due to the growing water demand
and increasing scarcity of water sources. As a result, the recycling of industrial wastewaters is
becoming an increasing need and stricter discharge standards continue to be introduced worldwide.
Therefore, the efficient removal of pollutants from wastewaters arises as an important area of
research.
The selection of a wastewater treatment among chemical, biological, and catalytic methods is
related to the toxicities and concentrations of the pollutants in the waste stream. The pollutant’s
concentrations should be high for chemical destruction methods that are thermally self-sufficient,
whereas bioprocesses are suitable for nontoxic pollutants at low concentrations (Guo and
Al-Dahhan, 2003). Chemical treatments such as flocculation, precipitation, adsorption on acti-
vated carbon, air stripping, reverse osmosis or ion-exchange do not convert pollutants, but rather
transfer them from a diluted to a concentrated stream, requiring a post-treatment (Centi and
Perathoner, 2005; Liotta et al., 2009). Thermal treatments present many drawbacks, such as severe
operation conditions and considerable emission of other hazardous compounds (Liotta et al.,
2009). Conventional biological processes represent an environmentally friendly way of treatment
with reasonable costs, however, they are not adequate to treat non-biodegradable wastewaters,
usually require a long residence time for microorganisms to degrade the pollutants and the dis-
posal of sludge formed during biological treatment can pose environmental problems associated
to additional costs (Guo and Al-Dahhan, 2003; Liotta et al., 2009).
Nowadays, advanced oxidation processes (AOPs) are a promising alternative, with the common
trend that they generate oxygen-based radicals in sufficient quantities to be able to oxidize the
majority of the complex chemicals present in the aqueous effluent (Gogate and Pandit, 2004).
Contaminants are oxidized through different reagents: ozone, hydrogen peroxide, oxygen and air
or their combination. These procedures may also be combined with UV radiation (Liotta et al.,
2009) and enhanced by catalytic materials. To choose the most appropriate technology, some
aspects, such as the concentration and nature of the pollutants and the volume of wastewater,
must be considered (Liotta et al., 2009).
Among AOPs, Fenton processes (using reaction between Fe ions and hydrogen peroxide, i.e.
Fenton’s reagent) have emerged as a viable alternative for the wastewater treatments of medium-
high total organic carbon concentrations (Liotta et al., 2009). Hydrogen peroxide does not form
any harmful by-products, and it is a non-toxic and ecological reactant. Moreover, although hydro-
gen peroxide is a relatively costly reactant, the peroxide oxidation compares very favorably to
processes that use gaseous oxygen or ozone. The lack of a gas/liquid boundary removes mass-
transfer limitations and the hydrogen peroxide acts as a free-radical initiator, providing hydroxyl
(HO) radicals that promote the degradation of organics. This allows lowering residence times and
enables conversion under milder conditions (Liotta et al., 2009). Furthermore, operating costs
are compensated by the lower fixed capital cost with respect to ozonation and wet air oxidation
(Centi et al., 2000).
81
82 N. Inchaurrondo, J. Font & P. Haure

The main reactions involved in the Fenton process are as follows (Pignatello et al., 2007):
H2 O2 + Fe2+ → Fe3+ + HO− + HO• (5.1)
+
H2 O2 + Fe 3+
→ Fe2+
+H + HO•2 (5.2)
• −
HO + Fe 2+
→ Fe 3+
+ HO (5.3)

HO + H2 O2 → H2 O + HO•2 (5.4)

HO + organics → products (5.5)

The mechanism of this reagent has not been fully explained because of the variety of Fe(II)
and Fe(III) complexes, numerous radical intermediate products and their consecutive reactions.
One advantage of the Fenton process is that no energy input is necessary to activate hydrogen
peroxide, making the reaction possible at atmospheric pressure and at room temperature. Fur-
thermore, this method requires relatively short reaction times and uses easy-to-handle reagents
(Gogate and Pandit, 2003). However, the traditional Fenton reaction needs a tight pH control
(between 3 and 4); requires further separation of the catalyst to prevent secondary water pollution
and it is strongly dependent on the presence of radical scavengers. Furthermore, the complexa-
tion of the cations by reaction products such as oxalic acid or inorganic ions such as phosphate,
may lead to a progressive lowering of the reaction rate and decrease of the H2 O2 consumption
efficiency (Centi and Perathoner, 2004; Fathima et al., 2007; Gogate and Pandit, 2003).
For these reasons, there has been a considerable interest in the development of heterogeneous
catalysts for the oxidation of wastewater streams. Recently, a great number of materials containing
iron or copper as precursors supported/intercalated on/in oxides, clays, zeolite, active carbon and
polymers have been proposed as active catalysts for Fenton-type reactions, to remove organic
compounds (Barrault et al., 2000; Bautista et al., 2011; Castro et al., 2010; Chahbane et al.,
2007; Crowther and Larachi, 2003; Dantas et al., 2006; Garrido-Ramírez et al., 2010; Guibal,
2005; Liou et al., 2005; 2009; Massa et al., 2011; Melero et al., 2006; Valkaj et al., 2007).
The heterogeneous catalytic system results less sensitive to the pH and more efficient in the
TOC abatement than the homogeneous one. The main difference between homogeneous and
heterogeneous systems is not the formation of different active oxygen species from H2 O2 , but
the ability of the heterogeneous catalyst to adsorb onto its surface the organic and/or the reaction
intermediate products, favoring then their reaction with oxygen species formed by H2 O2 activation
(Caudo et al., 2006).
These catalysts exhibit the advantages of a heterogeneously catalyzed process and show rela-
tively higher oxidation efficiency as well as a lower sensitivity to pH compared to homogeneous
catalysts at the same reaction conditions (Valkaj et al., 2007). Unlike the homogeneous systems,
solid catalysts can be recuperated by means of a simple separation operation and reused in next
runs. However, most of them cannot be used due to its lack of stability in aqueous media, because
of the leaching of the active elements or/and the support.
In this context, the preparation of stable materials for heterogeneous Fenton-type processes
having a good catalytic activity in a wide pH range and with negligible leaching of the transition
metal, need to be addressed.
Recent studies have focused on shifting from petrochemical-based feed-stocks to biological
materials to create high-performance and environmentally friendly catalysts. Polysaccharides
present many advantages that may stimulate their use as polymeric supports for catalysis: (i) they
are present in enormous quantity on earth, (ii) they have many functionalities that can be used
readily for the anchoring of organometallic species, (iii) they contain many stereogenic centers,
and (iv) they are chemically stable but biodegradable (Chtchigrovsky et al., 2009).
Among biopolymers chitin and chitosan are recommended as suitable functional materials
because these natural polymers have excellent properties such as biocompatibility, biodegrad-
ability, non-toxicity, adsorption properties, etc. (Viswanathana and Meenakshi, 2010). Chitin is a
linear chain consisting of N-acetyl-D-glucosamine (2-acetamido-2-deoxy-β-D-gluconopyranose)
joined together by β(1→4) linkage. It is found in abundance in exoskeletons of insects, shells
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 83

of crustaceans and fungal cell wall. It is the second most common polysaccharide occurring in
nature after cellulose.
Chitosan, a linear binary heteropolysaccharide, is composed of β-1,4-linked glucosamine
(GlcN) with various degrees of N-acetylation of GlcN residues. It is prepared by alkaline
N-deacetylation of chitin using concentrated sodium hydroxide (NaOH) solutions at high tem-
perature, for a long period of time and it is produced when the degree of deacetylation (DD)
is greater than 50%. This substance is only soluble in acidic aqueous solutions and insoluble in
water and alkaline solutions. When dissolved, the amino groups (–NH2 ) of the glucosamine are
protonated to –NH+ 3 . In mildly acidic solutions, this biopolymer can be readily cast into beads,
films, and fibers, allowing for great flexibility in manipulating and leading to a unique potential
as a catalyst (Macquarrie and Hardy, 2005).
Chitosan presents diverse properties enabling its use as a catalyst support (Guibal, 2005):

• It is characterized by a high nitrogen content, which explains, in turn, its ability to concentrate
metal ions and even neutral atoms of different metals via a variety of mechanisms such as ion
exchange or chelate formation, depending on the metal and pH of the solution.
• Easy modification and even possibility to use it without preliminary modification.
• High metal dispersion on the surface of a chitosan support.
• Rather high thermal stability, durability.
• Easy recovery of valuable components from the catalysts by incineration and metal extraction.

Its use as a catalyst support agrees with some of the “green chemistry” principles such as
the employment of alternative feedstock, more innocuous and renewable; the use of alternative
reaction conditions and the design of eco-compatible chemicals (less toxic than current alternatives
or inherently safer with regard to accident potential) (Centi and Perathoner, 2003). However, pure
chitosan has some disadvantages such as unsatisfactory mechanical properties, severe shrinkage,
deformation after drying, solubility under acidic conditions, and compressibility at high operating
pressure. Several methods to overcome these disadvantages have been performed such as coating
or impregnating chitosan on rigid porous materials. Coating chitosan as a thin layer onto an
immobilization support increases the accessibility of its binding sites, improves the mechanical
stability and reduces the amount of biopolymer needed (Futalan et al. 2011; Ngah et al., 2011;
Tsvetkova et al., 2007; Wan et al., 2005).
Chitosan has been immobilized in different supports such as bentonite, montmorillonite, perlite,
alumina, etc., and used as an adsorbent for several purposes like the removal of heavy metals
(Boddu et al., 2008; Chang et al., 2006; Dalida et al., 2011; Gandhi et al., 2010; Hasan et al.,
2008; Popuri et al., 2009) or dyes (Lee et al., 2009; Ngah et al., 2010; Wang and Wang, 2007;
Zhu et al., 2010). Moreover, chitosan has been use as a catalyst support for a variety of reactions:
oxidation of catechol to n-quinone (Yang and Vigee, 1991); reduction of 4-nitrophenol (Wei
et al., 2010); cyclopropanation of olefins (Wang et al., 2003); chemical fixation of carbon
dioxide to cyclic carbonate (Xiao et al., 2005); asymmetric hydrogenation of ketones (Wei et al.,
2004); oxidation of catecholamines (Chiessi and Pispisa, 1994); aldol and Knoevenagel reactions
(Kuhbeck et al., 2011); nitroaniline degradation (Vincent, 2004); allylic substitution reactions
(Beadoux et al., 2007); reduction of methylene blue (Rezende et al., 2010); carbonylation to
esters of Naproxen (Zhang et al., 2003); hydrogenation of furfuryl alcohol to tetrahydrofurfuryl
alcohol (Wu et al., 2007); hydrogenation for nitrobenzene (Huang et al., 2001); reduction of
chromate (Vincent and Guibal, 2002); Suzuki and Heck reactions (Hardi et al., 2004); synthesis
of Jasminaldehyde (Sudheesh et al., 2010); Huisgen Cycloaddition (Chtchigrovsky et al., 2009);
hydrolysis of phosphodiester (Chan and Cheng, 2009).
However, only few publications employed chitosan as a catalyst support for Fenton reactions.
Castro et al. (2009) studied different supported Cu(II) polymer catalysts (PVP2, PVP25, or chi-
tosan beads) for the catalytic oxidation of phenol at 30◦ C, at atmospheric pressure using air
or H2 O2 as oxidants. Sulakova et al. (2007) prepared Cu(II)/chitosan complexes for degrada-
tion of five model azo textile dyes in aqueous solution with hydrogen peroxide. Kucherov et al.
84 N. Inchaurrondo, J. Font & P. Haure

(2003) prepared series of heterogenized copper complexes by either coprecipitation or adsorp-


tion of Cu(II) on the bulk chitosan and composite supports (egg-shell type chitosan/SiO2 and
chitosan/MCM-41systems). Shen et al. (2010) prepared a CoTSPc@chitosan catalyst, by immo-
bilizing covalently water soluble cobalt (II) tetra-sulfophthalocyanine onto adsorbent chitosan
microspheres and used it in the peroxidation of C. I. Acid Red 73. Zubieta et al. (2008) evalu-
ated the removal of aqueous Methylene Blue (MB) and Benzopurpurin (BP) using TiO2 -chitosan
microporous materials. Lee et al. (2010) investigated the degradation of trichloroethylene (TCE),
at neutral pH, with a catalyst of Fe(II) chelated by cross-linked chitosan.
The objective of the present study is to synthesize and characterize a new Cu/chitosan composite
catalyst obtained by immobilizing Cu-chitosan complexes onto γ-alumina (γ-Al2 O3 ). The catalyst
was used in the Catalytic Wet Peroxide Oxidation (CWPO) of phenol in a fixed bed recirculating
reactor, at atmospheric pressure and moderate temperature of 50◦ C. Phenolic compounds removal
is a very active research field due to occurrence and the toxicity of phenolic pollutants in industrial
wastewaters and therefore phenol is often used as a model compound for Fenton studies (Barrault
et al., 2000; Crowther and Larachi, 2003; Inchaurrondo et al., 2012a; 2012b; Liotta et al., 2009;
Massa et al., 2011; Valkaj et al., 2007; Zazo et al., 2011).
The term catalyst immobilization or heterogeneization can be defined as “the transformation of
a soluble catalyst into a heterogeneous one, which is able to be separated from the reaction mixture
and preferably be reused for multiple times”. Several methods have been employed for linking
metal complexes to solids supports, either onto external surface or into the interior pores such as
formation of covalent bonds or noncovalent interactions (physisorption, electrostatic interactions,
H-bonding). Following green chemistry principles, physisorption was the method selected in this
contribution.

5.2 EXPERIMENTAL

5.2.1 Catalyst preparation and characterization


The Al2 O3 /chitosan-Cu(II) catalyst was prepared by coprecipitation of the Cu-chitosan complex
onto γ-Al2 O3 spheres. The Cu-chitosan complex was obtained by dissolving 4.5 g of chitosan
(medium molecular weight chitosan, 75–85% deacetylation, from Sigma–Aldrich) and 8.05 g
of CuCl2 ·2H2 O in 300 mL of HCl (0.1 mol L−1 ). The mixture was stirred until the formation
of a blue clear solution. Prior to the impregnation procedure, the γ-Al2 O3 (kindly provided by
Sasol, median particle diameter of 4.76 mm and total pore volume of 0.78 mL g−1 ) was washed
several times with distilled water. The sample was dried for 24 h at 50◦ C; it was then added into
the chitosan-Cu2+ solution in an Erlenmeyer of 250 mL and stirred for 30 min. The solid was
filtered and rinsed with distilled water several times before being contacted with a 0.1 mol L−1
NaOH solution for about 15 min, which neutralized the HCl within the chitosan gel and thereby
coagulated and deposited the Cu-chitosan complex on the γ-Al2 O3 surface. The sample was
filtered and washed again with distilled water until the filtrate was neutral, in order to remove
any NaOH. The particles were dried at 25◦ C for 48 h and then at 50◦ C for 24 h.
The copper content of the catalyst was determined by soaking the sample in 65% HNO3 solution
and then analyzing the copper concentration using Atomic Absorption Spectroscopy (AAS) in
an ANALYST 300 Perkin-Elmer Spectrophotometer. Copper leached from the catalyst during
oxidation tests was also measured by AAS. The specific surface area (BET) was measured with a
Micromeritics ASAP 2000 instrument using N2 adsorption at 77 K. Thermogravimetric analysis
(TGA) was performed to confirm the existence of the complex Cu-Chitosan onto the γ-Al2 O3 .
Samples of 18 mg approximately were taken for this purpose. Data were captured using a TGA
(Shimadzu, model TGA-50) instrument with a constant airflow of 20 mL min−1. The heating rate
applied was 10◦ C min−1 and the temperature was increased from 25 to 800◦ C. However, the
temperature was first kept constant at 110◦ C for 30 min in order to assure the complete removal
of water contained in the sample. Table 5.1 presents catalyst characterization results.
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 85

Table 5.1. Characterization of the freshAl2 O3 /chitosan-


Cu(II) catalyst.

Copper content (%) 4.20 ± 0.08


BET area (m2 g−1 ) 167 ± 10
Chitosan content (wt%.) 2.60 ± 0.38
Particle diameter (mm) 4.76

Figure 5.1. Set up of the trickle bed reactor used in this study.

5.2.2 Fenton like oxidation of phenol aqueous solutions


5.2.2.1 Reaction set-up
The CWPO was performed in a fixed bed-recirculating reactor packed with the homemade
Cu-chitosan/Al2 O3 catalyst. This set-up presents some advantages over the continuous reactor. It
is a smaller device that uses fewer amounts of catalyst and reactants and, although the one-pass
conversion is lower (differential reactor), much higher conversions can be achieved through pro-
longed exposure to the catalyst within the fixed bed. Additionally, as one-pass conversion does
not require to be large, there is no need of a significant liquid residence time in the reactor and this
allows extension of the examined liquid velocity range, using a reasonable amount of catalyst;
consequently, conditions attained at high and low velocities can be studied without changing the
set-up. An image of the reactor is shown in Figure 5.1.
A load of 41 g of the Cu(II)-chitosan/γ-Al2 O3 catalyst was used in the CWPO of a phenol
solution of 1 g L−1 (Panreac, 99%) in a recirculating jacketed packed bed glass reactor (60 cm
long and 2.35 cm of internal diameter) operated at 50◦ C and 1 atm in cocurrent downwards
operation. In this reactor, the ratio load of catalyst to load of liquid is high and this enables a rapid
mineralization of the intermediates generated during phenol oxidation, which helps preserving the
catalyst stability. In addition, it allows minimizing the contribution of secondary homogeneous
reactions and this higher concentration of active sites may have a beneficial effect on oxidation
performance due to the increasing amount of active sites for H2 O2 decomposition and for organic
86 N. Inchaurrondo, J. Font & P. Haure

compounds adsorption (Ramírez et al., 2007). The increase of copper sites should induce a
better use of oxidant for the promotion of free radicals, avoiding its thermal decomposition into
molecular oxygen and water (Sotelo et al., 2004).
Air was introduced at the top of the reactor at a flow rate of 50 mL min−1 . The liquid feed
consisted of 600 mL of a phenol solution, which was kept in a flask inside a thermostatic bath,
and recirculated at a flow rate of 45 mL min−1 through the column. A dose of 14.65 mL of H2 O2
(30 wt%) was added into the phenol solution flask to initiate the reaction. Once it was consumed,
a second dose of 14.65 mL was added. The total amount of added oxidant was 2.57 times the
stoichiometric requirement.
Five tests of 4 h each, at identical conditions, were performed, using the same load of catalyst
but with a fresh phenol solution each time. Catalyst leaching was evaluated over a total period
of 20 h. Phenol, TOC and H2 O2 concentrations and pH values were measured along the reaction
time. Experiments in the TBR (Trickle Bed Reactor) were repeated twice, starting from a fresh
catalyst bed but keeping other conditions constant. For simplicity, only the concentrations and
pH attained at the end of each run were measured. The error bars in Figures 5.7 and 5.9 represent
the difference between the two measurements. Experiments performed in the slurry reactor were
repeated at least three times and results are presented with the corresponding standard deviation.

5.2.3 Analytical methods


During CWPO tests, liquid samples were taken at different time intervals and immediately
analyzed. Phenol was detected and measured by a standard colorimetric method (Clesceri and
Greenberg, 1998). Hydrogen peroxide was detected by an iodometric titration method (Clesceri
and Greenberg, 1998).
Total Organic Carbon (TOC) values were obtained by a TOC analyzer (Analytic Jena, model
NC 2100). Samples were acidified with 50 mL HCl 2 mol L−1 , then were bubbled with synthetic
air for 3 min to eliminate the inorganic carbon content and finally injected. Phenol, TOC and
H2 O2 conversions were evaluated at different times up to 240 min. Data shown here represent
an average of at least three experiments, which showed satisfactory repeatability. Identification
of intermediates was carried out in a HPLC (Agilent Technologies, model 1100) with a C18
reverse phase column (Agilent Technologies, Hypersil ODS). The analyses were performed using
a mobile phase with a gradient mixture of methanol and ultrapure water (Milli-Q water, Millipore)
from 0/100 v/v to 40/60 v/v. The flow rate increased from 0.6 at the fifth minute to 4.0 mL min−1
at the seventh minute. The pH of the water was previously adjusted at 4.41 with sulfuric acid. A
diode array detector was used for detecting the organic compounds analyzed at 254 nm or 210 nm,
depending on the compound to be identified.

5.3 RESULTS AND DISCUSSION

5.3.1 Blank experiment


A preliminary blank experiment to discriminate the effect of adsorption was performed with the
γ-Al2 O3 support and in absence of oxidant. Without copper, phenol removal took place essentially
by adsorption, with phenol and TOC conversions of around 17%.

5.3.2 Activity and stability tests


The catalyst was used in five tests of 4 h each, under the same conditions, using a fresh phenol
solution each time. During the first test, a TOC conversion of 90% was obtained. Phenol and
H2 O2 were completely consumed. The H2 O2 was rapidly decomposed into radicals that allowed
a fast mineralization of the carboxylic acids generated during the oxidation of phenol. Phenol
and TOC conversions against time are presented in Figure 5.2, and H2 O2 concentration and pH
evolution in Figure 5.3.
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 87

Figure 5.2. Evolution of phenol and TOC conversions over time. 600 mL of 1 g L−1 phenol; 41 g of 4.2%
Cu catalyst bed; phenol solution flow: 45 mL min−1 ; temperature: 323 K; initial dose of H2 O2 :
4.3 times the stoichiometric value.

Figure 5.3. Evolution of pH and H2 O2 concentration over time. 600 mL of 1 g L−1 phenol; 41 g of 4.2%
Cu catalyst bed; phenol solution flow: 45 mL min−1 ; temperature: 323 K; initial dose of H2 O2 :
4.3 times the stoichiometric value.

The HPLC identification of intermediates indicated the presence of a low concentration of aro-
matics (hydroquinone, catechol, p-benzoquinone), which were finally oxidized into biodegradable
carboxylic acids (mostly succinic, maleic, malonic and oxalic acids). The distribution of inter-
mediates is presented in Figure 5.4 and the percentage of phenol, identified and unidentified
compounds are shown in Figure 5.5.
Initially, phenol is converted to hydroquinone and catechol. As catechol is the main primary
oxidation product, hydroxylation seems to take place predominantly at the ortho position. Next,
these compounds are oxidized to o- and p-benzoquinone. These colored intermediates are much
more toxic than phenol itself and thus their presence even at low concentration would mean high
values of ecotoxicity (Suárez-Ojeda et al., 2007; Zazo et al., 2005). Then, ring-opening of the aro-
matic intermediates leads to the formation of organic acids. Fumaric, maleic, malonic, succinic
and oxalic acids were the identified products at this stage. During the experiment, the color of
the reaction system turned from transparent to light pink due to the presence of p-benzoquinone
and became again clear as the reaction progressed. The addition of a second oxidant dose at
88 N. Inchaurrondo, J. Font & P. Haure

Figure 5.4. Intermediates distribution against time, first test, 600 mL of 1 g L−1 phenol; 41 g of 4.2% Cu
catalyst bed; phenol solution flow: 45 mL min−1 ; temperature: 323 K; initial dose of H2 O2 : 4.3
times the stoichiometric value.

Figure 5.5. Percentage of phenol, identified and unidentified products in the intermediates distribution
against reaction time, during the first test.

120 min resulted in a substantial modification of intermediate product distribution. In fact, phenol
decomposition was observed at a rate even higher than that initially observed. Consequently, the
concentration of most identified intermediates (aromatics and malonic) increased up to 135 min
approximately. Fumaric acid concentration decreases with no significant changes in its decom-
position rate and succinic acid concentration decreases up to 150 min and then increases up to
195 min. At 240 min reaction time, the concentration of all identified by products was negligible.
As it is shown in Figure 5.5 in this HPLC preliminary analysis it was not feasible to identify
a high percentage of reaction intermediates (possibly dimmers and high condensation products),
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 89

Figure 5.6. pH evolution against time for each test. 600 mL of 1 g L−1 phenol; 41 g of 4.2% Cu catalyst
bed; phenol solution flow rate: 45 mL min−1 ; temperature: 323 K; initial dose of H2 O2 : 4.3
times the stoichiometric value.

Figure 5.7. Final TOC and phenol conversions and residual peroxide concentrations at the end of each
run. 600 mL of 1 g L−1 phenol; 41 g of 4.2% Cu catalyst bed; phenol solution flow rate:
45 mL min−1 ; temperature: 323 K; initial dose of H2 O2 : 4.3 times the stoichiometric value.

a situation commonly encountered by other researchers (Villota et al., 2007; Zazo et al., 2005;
2009).
The high degree of mineralization and low accumulation of acids was proved by the pH evo-
lution, which stayed above 6 during most of the reaction, with a minimum of 5.3. Nevertheless,
it was suspected that some carboxylic acids were adsorbed by the catalyst. During the first four
hours of use, no copper was detected in the liquid phase; therefore, metal leaching was negligible.
Catalyst stability was addressed using the same load in four identical consecutive runs of four
hours each. Catalyst was not pretreated between tests. In each experiment, the equipment was
loaded with fresh phenol and hydrogen peroxide solutions. Runs were performed at identical
operating conditions, except that, due to the accumulation of carboxylic acids, the initial pH
values were progressively lower (up to 5.5). Furthermore, the pH evolutions also presented lower
pH values, as seen in Figure 5.6.
The final TOC and phenol conversions and the residual H2 O2 concentration at the end of
each test are presented in Figure 5.7. During the following runs, complete phenol conversion
was obtained at the end of each reaction. However, the H2 O2 decomposition and final TOC
90 N. Inchaurrondo, J. Font & P. Haure

Figure 5.8. Color evolution of the reaction media in the fifth test (time interval: 0, 15, 30, 60, 90, 120, 180,
240 min).

Figure 5.9. Copper leached from the catalyst at the end of each reaction.

conversion decrease with use. The final concentration of oxidant increases due to a lower rate of
radical generation, which causes the TOC conversion to drop. As a result, the final TOC conversion
decreased from an 89.7% in the first test up to a 69% after 20 h of usage (20% reduction).
Variations in the effluent color and pH evolution evidenced a different product distribution.
The color of the supernatant at the end of the following reactions turn brown due to the presence
of benzoquinones and the pH reached lower values due to a higher accumulation of carboxylic
acids. As an example, color evolution of the reaction media with a catalyst used 16 h is presented
in Figure 5.8. Furthermore, catalyst color changed from light blue to light brown as time elapses,
due to the presence of adsorbed intermediates.
The amount of copper leached from the catalyst at the end of each reaction and present in the
liquid phase was evaluated by AA experiments. Results (as % leaching and ppm) are presented
Figure 5.9. Taking into account these measurements, it was possible to evaluate the copper content
of the catalyst at any time, subtracting the total leached copper from the initial load. Using this
approach, the copper load decreased from 4.2 to 1% after 20 h of usage.
The homogeneous contribution should not be significant once compared to the heterogeneous
phenomena, as the catalyst load employed in this experiment is quite high, compared to the amount
of copper leached. Furthermore, the homogeneous catalytic contribution due to the leached metal
ion in a batch reaction varies with time as well as reaction conditions (Njiribeako et al., 1978).
The importance of the heterogeneous phenomena is evidenced in the first run, where no leaching
takes place and the final conversions are the most significant. Moreover, the copper leached can
be directly sequestered by organic acids and diacids forming copper species often having lower
activity (Caudo et al., 2007).
The results presented in Figures 5.7 and 5.9 are a clear signal of reduced performance,
which becomes more evident after 8 h of use. Several aspects could contribute to the observed
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 91

Table 5.2. Comparison of fresh and used Al2 O3 /chitosan-Cu(II) catalyst performances
in a slurry reactor.

Catalyst sample TOC conversion (%) Phenol conversion (%)

Fresh 56 ± 3 100
Used 57 ± 5 100

Table 5.3. Effect of initial pH on reaction outcomes, with fresh catalyst in a slurry reactor.

Initial pH TOC conversion (%) Phenol conversion (%) H2 O2 conversion (%)

5 75.5 ± 5 100 79 ± 4
7 74 ± 5.6 100 75 ± 3.6

results: metal leaching, adsorption of intermediate products and lower initial pH conditions
(these three factors directly affect catalyst activity) and even poor liquid distribution (leading to
decreasing catalyst utilization due to localized catalyst deactivation). Therefore, additional exper-
iments were carried out to independently assess the contribution of each aspect on the overall
performance.

5.3.3 Deactivation phenomena


To discriminate the effects of leaching, adsorption of intermediate products and initial pH, on
catalyst performance, additional tests were performed in a slurry reactor, in which wetting is
complete.

5.3.4 Effect of intermediate products adsorption


To evaluate the effect of adsorption of intermediates, phenol oxidation tests were done comparing
the performance of used and fresh catalyst with the same initial amount of metal present, at
identical initial pH conditions and with complete wetting (slurry reactor). Experiments were
carried out in a 250 mL stoppered glass reactor equipped with a condenser, thermocouple and a
pH electrode, provided with vigorous agitation and in contact with air at atmospheric pressure.
Tests were performed at 50◦ C for 1 hour. Amounts of 0.50 g of used and 0.42 g of fresh catalyst
were employed, which is equivalent to 29.4 ppm of copper in the solution, therefore experiments
were performed using the same initial load of copper. Catalysts were contacted with 170 mL of
0.1 g L−1 of phenol aqueous solution under continuous stirring. An excess of hydrogen peroxide
was added to the system (190 mmol L−1 ) and the reaction started.
The TOC and phenol conversions achieved at 60 min are presented in Table 5.2.
According to these results, the adsorption of intermediates has no effect on TOC and phenol
conversions.

5.3.5 Initial pH effect


Tests were performed in the slurry reactor using 2 g of fresh catalyst, at 50◦ C, with 170 mL of
1 g L−1 phenol solution and a hydrogen peroxide initial concentration of 190 mmol L−1 (4.3 the
stoichiometric requirement). The first test was carried out at pH = 7, a value close to the usual
initial pH of the phenol solution. The second test was performed with an initial pH of 5, adjusted
by adding a few drops of HCl. The final TOC, phenol and H2 O2 conversions achieved at 240 min
are presented in Table 5.3.
The initial pH has little effect on the reaction evolution, at least in this pH range.
92 N. Inchaurrondo, J. Font & P. Haure

Table 5.4. Effect of catalyst load on reaction outcomes, with fresh catalyst, in a slurry
reactor.
Copper load TOC conversion Phenol conversion H2 O2 conversion
(g L−1 ) (%) (%) (%)

19.3 89.5 ± 2.34 100 ± 0.84 96 ± 2.47


9.7 75.8 ± 5.6 98 ± 0.53 82 ± 3.57
4.8 60 ± 4.3 96 ± 0.76 66 ± 2.34

Figure 5.10. Phenol conversion evolution against time for each phenol flow rate (45 mL min−1 and
180 mL min−1 ).

5.3.6 Copper load effect


As it was found that the initial pH and adsorbed intermediates have no significant effect on the final
TOC conversion, the effect of copper load was analyzed in experiments performed in the slurry
reactor using different amounts of fresh catalyst, at 50◦ C, with 170 mL of 1 g L−1 phenol solution,
a hydrogen peroxide initial concentration of 190 mmol L−1 (4.3 the stoichiometric requirement)
and initial pH = 7. Results are presented in Table 5.4.
Increasing the copper content by a factor of almost 4 (while the other conditions remain
constant) renders an enhancement of TOC conversion of 30%. Accordingly, it is uncertain to
strictly attribute the 20% TOC reduction depicted in Figure 5.7 (from 89.7 to 69%) to the 20%
loss of copper measured.
However, the liquid-catalyst contact (or wetting efficiency, f ) may not be complete along the
reactor, leading to some unused catalyst fraction ( f = 0.55) (Baussaron et al., 2007). Therefore,
the leaching effect could be more significant in the fraction of the trickle bed that is actually used
(wetted fraction) during the reaction.

5.3.7 Liquid flow rate effect


The effect of wetting on performance was consequently explored. A test after the fifth run was
carried out using a phenol flow rate of 180 mL min−1 ( f = 0.8). A TOC conversion of 83.5% was
obtained after 240 min. In spite of the lower copper content of the used catalyst, the adsorbed
intermediate products and the lower initial pH, the TOC conversion attained at a higher liquid
flow rate was 20% larger than that obtained at the end of the fifth run (69%), but using a small
liquid flow rate.
Additional experiments were performed in order to specifically assess the effect of phenol
flow rate. The tests were carried out in the fixed bed recirculating reactor packed with identical
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 93

Figure 5.11. TOC conversion evolution against time for each phenol flow rate (45 mL min−1 and 180 mL
min−1 ).

Figure 5.12. H2 O2 concentration evolution against time for each phenol flow rate (45 mL min−1 and
180 mL min−1 ).

catalyst loads previously used during 4 h. Two different phenol flow rates were investigated: 45
and 180 mL min−1. Figures 5.10 and 5.11, respectively, present phenol and TOC conversions
against time, for both liquid flow rates. Figures. 5.12 and 5.13 present respectively the H2 O2
concentration and pH evolution, for each flow.
With a flow rate of 180 mL min−1 , the initial phenol, H2 O2 and TOC conversion rates resulted
higher than those achieved with the lower flow rate. Phenol was completely oxidized in approxi-
mately 100 min and a TOC reduction of 80% was obtained in 120 min with a final conversion of
90% at the end of the reaction. When the flow rate is increased, the hydrodynamic conditions are
modified and the contact between the liquid and the catalyst or wetting efficiency is improved
from 0.55 to 0.8 approximately. This favors the generation of radicals and also the termination
reactions over the catalyst surface.
Just like in the previous tests, and additional H2 O2 dose of 14.65 mL was added at 120 min. This
extra oxidant amount has nil impact on the already high phenol and TOC conversions, achieved
at the highest liquid flow rate but increases phenol and TOC conversions obtained working with
a 45 mL min−1 flow rate. For both experiments, the additional oxidant dose increases pH values
due to the mineralization of carboxylic acids. The measured leaching of copper was the same
under these different conditions. In both tests, the leaching resulted less than 2%.
94 N. Inchaurrondo, J. Font & P. Haure

Figure 5.13. pH evolution against time for each phenol flow rate (45 mL min−1 and 180 mL min−1 ).

5.4 CONCLUSIONS

The home-made Cu-chitosan/γ-Al2 O3 catalyst successfully oxidizes a phenol solution of 1 g L−1


in a laboratory recirculating packed bed glass reactor operated at 323 K and 1 atm, employing
gas and liquid flow rates of 50 and 45 mL min−1 respectively in cocurrent downwards operation.
Complete phenol and H2 O2 consumption and a TOC conversion of 90% were obtained after four
hours with a small excess of H2 O2 . No metal leaching was detected at this time.
Catalyst stability was addressed using the same load in four identical consecutive runs of four
hours each, during 20 h. A complete phenol conversion was maintained but the TOC conversion
decreased from 89.5% in the first run to 69% at the end of the last test. A final copper leaching
of 20% was estimated. In spite of this loss, the system still maintained a high copper load;
therefore, it was uncertain to attribute the deactivation only to the leaching of the active metal.
The adsorption of intermediate products and drop of the initial pH were proven not to be related
to the deactivation process. However, the incomplete liquid-catalyst contact led to some fraction
of unused catalyst, which accentuated the leaching effect in the fraction of the trickle bed that
was actually used. This was proven by increasing the phenol flow rate from 45 mL min−1 to
180 mL min−1 , in a test performed with a load of catalysts used during 20 h. The increase of the
liquid flow rate favored the effective wetting of the catalyst, promoting the generation of more
radicals and also the termination reactions over the catalyst surface. Consequently, the final TOC
conversion achieved at better wetting conditions (89%) was significantly higher than the value
obtained in the previous test (69%), with a lower flow rate.
The effect the phenol flow rate was further studied. With a higher flow rate, the initial phenol,
TOC and H2 O2 rates increased in comparison with results obtained at the lower flow rate. With
the highest flow rate, complete conversion of phenol and 80% conversion of TOC were achieved,
in only 120 min, before adding the second dose of oxidant. According to these results, under
these conditions, it is possible to obtain a higher mineralization with a lower amount of oxidant
by increasing the phenol flow rate.
Finally, even though the catalyst proved to be active for phenol mineralization, copper leaching
was detected, which could be removed by precipitation, membrane filtration or adsorption before
feeding the effluent into a subsequent treatment step.

ACKNOWLEDGEMENTS

Financial support from CONICET, UNMdP, and ANPCyT (Argentina), AUIP, AECID (projects
A/9419/07 and A/017018/08), and the Spanish Ministerio de Ciencia e Innovación (project
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 95

CTM2008-03338) are gratefully acknowledged. We also want to express our gratitude to


C. Rodriguez, J. Cechini and H. Asencio for their technical support.

REFERENCES

Barrault, J., Tatibouët, J.M. & Papayannakos, N.: Catalytic wet peroxide oxidation of phenol over pillared
clays containing iron or copper species. C. R. Acad. Sci. Paris, Série IIc, Chimie: Chemistry 3 (2000)
pp. 777–783.
Baudoux, J., Perrigaud, K., Madec, P-J., Gaumont, A-C. & Dez, I.: Development of new SILP catalysts using
chitosan as support. Green Chem. 9 (2007), pp. 1346–1354.
Baussaron, L., Julcour-Lebigue, C., Boyer, C., Wilhelm, A.M. & Delmas, H.: Effect of partial wetting on
liquid/solid mass transfer in trickle bed reactors. Chem. Eng. Sci. 62:24 (2007), pp. 7020–7025.
Bautista, P., Mohedano, A.F., Casas, J. A., Zazo, J. A. & Rodriguez, J.J.: Highly stable Fe/γ -Al2 O3 catalyst
for catalytic wet peroxide oxidation. J. Chem. Technol. Biotechnol. 86 (2011), pp. 497–504.
Boddu, V.M., Abburi, K., Talbott, J.L., Smith, E.D., & Haasch, R.: Removal of arsenic (III) and arsenic (V)
from aqueous medium using chitosan-coated biosorbent. Water Res. 42 (2008), pp. 633–642.
Castro, I.U., Stüber, F., Fabregat, A., Font, J., Fortuny, A. & Bengoa, C.: Supported Cu(II) polymer catalysts
for aqueous phenol oxidation. J. Hazard. Mater. 163 (2009), pp. 809–815.
Castro, I.U., Sherrington, D.C., Fortuny, A., Fabregat, A., Stüber, F., Font, J. & Bengoa, C.: Synthesis of
polymer-supported copper complexes and their evaluation in catalytic phenol oxidation. Catal. Today 157
(2010), pp. 66–70.
Caudo, S., Centi, G., Genovese, C. & Perathoner, S.: Homogeneous versus heterogeneous catalytic reactions
to eliminate organics from waste water using H2 O2 . Top. Catal. 40 (2006), pp. 207–219.
Caudo, S., Centi, G., Genovese, C. & Perathoner, S.: Copper- and iron-pillared clay catalysts for the WHPCO
of model and real wastewater streams from olive oil milling production. Appl Catal B: Environ. 70 (2007),
pp. 437–446.
Centi, G. & Perathoner, S.: Catalysis and sustainable (green) chemistry. Catal. Today 77 (2003), pp. 287–297.
Centi, G., Perathoner, S., Torre, T. & Verduna, M. G.: Catalytic wet oxidation with H2 O2 of carboxylic acids
on homogeneous and heterogeneous Fenton-type catalysts. Catal. Today 55 (2000), pp. 61–69.
Centi, G. & Perathoner, S.: Use of solid catalysts in promoting water treatment and remediation technolo-
gies. In: J.J. Spivey (ed): Catalysis. The Royal Society of Chemistry, Cambridge, England, 2005, 18,
pp. 46–74.
Chahbane, N., Popescu, D.L., Mitchell, D.A., Chanda, A., Lenoir, D., Ryabov, A.D., Schramma, K.-W. &
Collins, T.J.: FeIII–TAML-catalyzed green oxidative degradation of the azo dye Orange II by H2 O2 and
organic peroxides: Products, toxicity, kinetics, and mechanisms. Green Chem. 9 (2007), pp. 49–57.
Chang, Y-C., Chang, S.W. & Chen, D.-H.: Magnetic chitosan nanoparticles: Studies on chitosan binding and
adsorption of Co(II) ions. React. Funct. Polym. 66 (2006), pp. 335–344.
Chang, Y-C. & Chen, D.-H.: Highly efficient hydrolysis of phosphodiester by a copper(II)-chelated chitosan
magnetic nanocarrier. React. Funct. Polym. 69 (2009), pp. 601–605.
Chiessi, E. & Pispisa, B.: Polymer-supported catalysis: Oxidation of catecholamines by Fe (III) and Cu (II)
complexes immobilized to chitosan. J. Mol. Catal. 87 (1994), pp. 177–194.
Clesceri, L.S., Greenberg, A.E., Eaton A.D. (eds): Standard methods for the examination of water and
wastewater. 20th ed. American Public Health Association, Washington, DC, 1998.
Chtchigrovsky, M., Primo, A., Gonzalez, P., Molvinger, K., Robitzer, M., Quignard F. & Taran, F.: Function-
alized chitosan as a green, recyclable, biopolymer-supported catalyst for the [3+2] Huisgen Cycloaddition.
Chem. Int. Ed. 48 (2009), pp. 5916–5920.
Crowther, N. & Larachi, F.: Iron-containing silicalites for phenol catalytic wet peroxidation. Appl. Catal. B
46 (2003), pp. 293–305.
Dalida, M.L.P., Mariano, A.F., Futalan, C.M., Kan, C.-C., Tsai, W.-C. & Wan, M.W.: Adsorptive removal of
Cu(II) from aqueous solutions using non-crosslinked and crosslinked chitosan-coated bentonite beads.
Desalination 275 (2011), pp. 154–159.
Dantas, T.L.P., Mendonca, V.P., José, H.J., Rodrigues, A.E. & Moreira, R.F.P.M.: Treatment of textile wastew-
ater by heterogeneous Fenton process using a new composite Fe2 O3 /carbon. Chem. Eng. J. 118 (2006),
pp. 77–82.
Fathima, N.N., Aravindhan, R., Rao, J.R. & Nair, B.U.: Dye house wastewater treatment through advanced
oxidation process using Cu-exchanged Y zeolite: A heterogeneous catalytic approach. Chemosphere 70
(2008), pp. 1146–1154.
96 N. Inchaurrondo, J. Font & P. Haure

Futalan, C.M., Kan, C.-C., Dalida, M.L., Hsien, K.J., Pascual, C. & Wan, M.-W.: Comparative and competi-
tive adsorption of copper, lead, and nickel using chitosan immobilized on bentonite. Carbohydr. Polym. 83
(2011), pp. 528–536.
Gandhi, M.R., Viswanathanc, N. & Meenakshi, S.: Preparation and application of alumina/chitosan
biocomposite. Int. J. Biol. Macromol. 47 (2010), pp. 146–154.
Gogate, P.R. & Pandit, A.B.: A review of imperative technologies for wastewater treatment I: Oxidation
technologies at ambient conditions. Adv. Environ. Res. 8 (2004), pp. 501–554.
Guibal, E.: Heterogeneous catalysis on chitosan-based materials: A review. Prog. Polym. Sci. 30 (2005),
pp. 71–109.
Guo, J. & Al-Dahhan, M.: Catalytic wet oxidation of phenol by hydrogen peroxide over pillared clay catalyst.
Eng. Chem. Res. 42 (2003), pp. 2450–2460.
Hasan, S., Ghosha, T.K., Viswanath, D.S. & Boddu, V.M.: Dispersion of chitosan on perlite for enhancement
of copper(II) adsorption capacity. J. Hazard. Mater. 152 (2008), pp. 826–837.
Hardy, J.J.E., Hubert, S., Macquarrie, D.J. & Wilson, A.J.: Chitosan-based heterogeneous catalysts for Suzuki
and Heck reactions. Green Chem. 6 (2004), pp. 53–56.
Huang, A., Liu, Y., Chen, L. & Hua, J.: Synthesis and property of nanosized palladium catalysts protected
by chitosan/silica. J. Appl. Polym. Sci. 85 (2002), pp. 989–994.
Inchaurrondo, N., Cechini, J., Font, J. & Haure, P.: Strategies for enhanced CWPO of phenol solutions. Appl
Catal B: Environ. 111–112 (2012a), pp. 641–648.
Inchaurrondo, N., Massa, P., Fenoglio, R., Font, J. & Haure, P.: Efficient catalytic wet peroxide oxidation
of phenol at moderate temperature using a high-load supported copper catalyst. Chem. Eng. J. 198–199
(2012b), pp. 426–434.
Kucherov, A.V., Kramareva, N.V., Finashina, E.D., Koklin A.E. & Kustov L.M.: Heterogenized redox catalysts
on the basis of the chitosan matrix 4. Copper complexes. J. Mol. Catal. A: Chem. 198 (2003), pp. 377–389.
Kühbeck, D., Saidulu, G., Rajender Reddy, K. & Díaz Díaz, D.: Critical assessment of the efficiency of
chitosan biohydrogel beads as recyclable and heterogeneous organocatalyst for C–C bond formation.
Green Chem. 14 (2012), pp. 378–392.
Lee, Y. & Lee, W.: Degradation of trichloroethylene by Fe(II) chelated with cross-linked chitosan in a
modified Fenton reaction. J. Hazard. Mater. 178 (2010), pp. 187–193.
Liotta, L.F., Gruttadauria, M., Di Carlo, G., Perrini, G. & Librando, V.: Heterogeneous catalytic degradation
of phenolic substrates: Catalysts activity. J. Hazard. Mater. 162 (2009), pp. 588–606.
Liou, R.-M., Chen, S.-H., Hung, M.-Y., Hsu, C.-S. & Lai J.-Y.: Fe (III) supported on resin as effective catalyst
for the heterogeneous oxidation of phenol in aqueous solution. Chemosphere 59 (2005), pp. 117–125.
Liou, R.-M. & Chen, S.-H.: CuO impregnated activated carbon for catalytic wet peroxide oxidation of phenol.
J. Hazard. Mater. 172 (2009), pp. 498–506.
Macquarrie, D.J. & Hardy, J.J.E.: Applications of functionalized chitosan in catalysis. Ind. Eng. Chem. Res.
44 (2005), pp. 8499–8520.
Massa, P.A., Ivorra, F., Haure,. P. & Fenoglio, R.J.: Catalytic wet peroxide oxidation of phenol solutions over
CuO/CeO2 systems. J. Hazard. Mater. 190 (2011), pp. 1068–1073.
Melero, J.A., Calleja, G., Martínez, F. & Molina, R.: Nanocomposite of crystalline Fe2 O3 and CuO particles
and mesostructured SBA-15 silica as an active catalyst for wet peroxide oxidation processes. Catal.
Commun. 7 (2006), pp. 478–483.
Ngah, W.S.W., Md. Ariff, N.F. & Hanafiah, M.A.K.M.: Preparation, characterization, and environmental
application of crosslinked chitosan-coated bentonite for tartrazine adsorption from aqueous solutions.
Water Air Soil Pollut. 206 (2010), pp. 225–236.
Ngah, W.S.W., Teong, L.C. & Hanafiaha, M.A.K.M.: Adsorption of dyes and heavy metal ions by chitosan
composites: A review. Carbohydr. Polym. 83 (2011), pp. 1446–1456.
Njiribeako A.I., Hudgins R.E. & Silveston P.L., Catalytic oxidation of phenol in aqueous solution over copper
oxide. Ind. Eng. Chem. Fundam. 17:3 (1978), pp. 234–234.
Pignatello, J.J., Oliveros, E. & MacKay, A.: Advanced oxidation processes for organic contaminant destruc-
tion base on the fenton reaction and related chemistry. Crit. Rev. Env. Sci. Technol. 36:1 (2007),
pp. 1–84.
Popuri, S.R., Vijaya, Y., Boddu, V.M. & Abburi, K.: Adsorptive removal of copper and nickel ions from water
using chitosan coated PVC beads. Bioresour. Technol. 100 (2009), pp. 194–199.
Rezende, T.S., Andrade, G.R.S., Barreto, L.S., Costa, Jr. N.B., Gimenez, I.F. & Almeida, L.E.: Facile
preparation of catalytically active gold nanoparticles on a thiolated chitosan. Mater. Lett. 64 (2010),
pp. 882–884.
Fenton-like oxidation of phenol with a Cu-chitosan/Al2 O3 catalyst 97

San Sebastián Martínez, N., Fíguls Fernández, J., Font Segura, X. & Sánchez Ferrer, A.: Pre-oxidation
of an extremely polluted industrial wastewater by the Fenton’s reagent. J. Hazard. Mater. B101 (2003),
pp. 315–322.
Shen, C., Song, S., Zang, L., Kang, X., Wen, Y., Liu, W. & Fu, L.: Efficient removal of dyes in water using
chitosan microsphere supported cobalt (II) tetrasulfophthalocyanine with H2 O2 . J. Hazard. Mater. 177
(2010), pp. 560–566.
Suarez-Ojeda, M.A., Guisáosla, A., Baeza, J.A., Fabregat, A., Stüber, F., Fortuna, A., Font, J. & Carrera,
J.: Integrated catalytic wet air oxidation and aerobic biological treatment in a municipal WWTP of a
high-strength o-cresol wastewater. Chemosphere 66 (2007), pp. 2096–2105.
Sudheesh, N., Sharma, S. K. & Shukla, R.S.: Chitosan as an eco-friendly solid base catalyst for the solvent-
free synthesis of jasminaldehyde. J. Mol. Catal. A: Chem. 321 (2010), pp. 77–82.
Suláková, R., Hrdina, R. & Soares, G.M.B.: Oxidation of azo textile soluble dyes with hydrogen peroxide in
the presence of Cu(II) chitosan heterogeneous catalysts. Dyes Pigm. 73 (2007), pp. 19–24.
Valkaj, K.M., Katovic, A. & Zrncevic, S.: Investigation of the catalytic wet peroxide oxidation of phenol over
different types of Cu/ZSM-5 catalyst. J. Hazard. Mater. 144 (2007), pp. 663–667.
Valkaj, K.M., Katovic, A. & Zrncevic, S.: Catalytic properties of Cu/13X zeolite based catalyst in catalytic
wet peroxide oxidation of phenol. Ind. Eng. Chem. Res. 50 (2011), pp. 4390–4397.
Villota, N., Mijangos, F., Varona, F. & Andrés, J.: Kinetic modelling of toxic compoundsgenerated during
phenol elimination in wastewaters. Int. J. Chem. Reactor Eng. 5 (2007) A63.
Vincent, T. & Guibal, E.: Chitosan-supported palladium catalyst. 4. Synthesis procedure. Ind. Eng. Chem.
Res. 41 (2002), pp. 5158–5164.
Vincent, T., Peirano, F. & Guibal, E.: Chitosan supported palladium catalyst. VI. Nitroaniline degradation. J.
Appl. Polym. Sci. 94 (2004), pp. 1634–1642.
Viswanathana, N. & Meenakshi, S.: Enriched fluoride sorption using alumina/chitosan composite. J. Hazard.
Mater. 178 (2010) pp. 226–232.
Wan, M.-W., Petrisor, I.G., Lai, H.-T., Kim, D. & Yen, T.F.: Copper adsorption through chitosan immobilized
on sand to demonstrate the feasibility for in situ soil decontamination. Carbohydr. Polym. 55 (2004),
pp. 249–254.
Wang, L. & Wang, A.: Adsorption characteristics of Congo Red onto the chitosan/montmorillonite
nanocomposite. J. Hazard. Mater. 147 (2007), pp. 979–985.
Wang, W., Sun, W. & Xia, C.: An easily recoverable and efficient catalyst for heterogeneous cyclopropanation
of olefins. J. Mol. Catal. A: Chem. 206 (2003) pp. 199–203.
Wei, D., Ye, Y., Jia, X., Yuan, C. & Qian, W.: Chitosan as an active support for assembly of metal nanoparticles
and application of the resultant bioconjugates in catalysis. Carbohydr. Res. 345 (2010), pp. 74–84.
Wei, W.-L., Hao, S.-J., Zhou, J., Huang, M.-Y. & Jiang, Y.-Y.: Catalytic behavior of silica-supported chitosan-
platinum-iron complex for asymmetric hydrogenation of ketones. Polym. Adv. Technol. 15 (2004),
pp. 287–290.
Wu, F-C., Tseng, R.-L. & Juang, R-S.: Kinetic modeling of liquid phase adsorption of reactive dyes and
metal ions on chitosan. Water Res. 35 (2001), pp. 613–618.
Xiao, L-F., Li, F-W. & Xia, C-G.: An easily recoverable and efficient natural biopolymer-supported zinc
chloride catalyst system for the chemical fixation of carbon dioxide to cyclic carbonate. Appl. Catal. A
279 (2005), pp. 125–129.
Yang, J. H. & Vigee, G.S.: A cobalt complex immobilized to chitosan. J. Inorg. Biochem. 41 (1991), pp. 7–16.
Zazo, J.A., Casas, J.A., Mohedano, A.F., Gilarranz, M.A. & Rodriguez, J.: Chemical pathway and kinetics
of phenol oxidation by Fentons reagent. Environ. Sci. Technol. 39 (2005), pp. 9295–9302.
Zazo, J.A., Casas, J.A., Mohedano, A.F. & Rodríguez, J.J.: Semicontinuous Fenton oxidation of phenol in
aqueous solution: A kinetic study. Water Res. 43 (2009), pp. 4063–4069.
Zazo J.A., Pliego G., Blasco S., Casas J.A. & Rodríguez J.J.: Intensification of the Fenton process by
increasing the temperature. Ind. Eng. Chem. Res. 50 (2011) pp. 866–870.
Zhang, J. & Xia, C.-G.: Natural biopolymer-supported bimetallic catalyst system for the carbonylation to
esters of Naproxen. J. Mol. Catal. A: Chem. 206 (2003), pp. 59–65.
Zhu, H.Y., Jiang, R. & Xiao, L.: Adsorption of an anionic dye by chitosan/kaolin/γ-Fe2 O3 composites. Appl
Clay Sci. 48 (2010), pp. 522–526.
Zubieta, C.E., Messina, P.V., Luengo, C., Dennehy, M., Pieroni, O. & Schulz, P.C.: Reactive dyes remotion
by porous TiO2 -chitosan materials. J. Hazard. Mater. 152 (2008), pp. 765–777.
This page intentionally left blank
CHAPTER 6

Degradation of a mixture of glyphosate and 2,4-D in water solution


employing the UV/H2 O2 process, including toxicity evaluation

Melisa Mariani, Roberto Romero, Alberto Cassano & Cristina Zalazar

6.1 INTRODUCTION

Glyphosate (N -phosphonomethyl glycine) is one of the herbicides most widely used throughout
the world due to its broad spectrum, low cost, and non-selectivity. It can be used for non-crop
land as well as for a great variety of crops.
Several factors have contributed to increase the agricultural use of glyphosate in the last decades:
price reductions, increase in supply related to patent expiration, and mainly the implementation
of genetically modified (GM) glyphosate resistants (GR) cultivars.
The countries with the highest production of soybean are USA, Brazil and Argentina. According
to 2010 statistics, they produce 90, 68 and 18 million tons of soybeans, respectively (FAO,
FAOSTAT). GR soybean is the major crop in Argentina and its production has been accompanied
by an even more important increase in the use of glyphosate. InArgentina, glyphosate use increased
from 1 million to 180 million liters between 1997 and 2007 (Binimelis et al., 2009).
The widespread use of this herbicide causes two important problems: water pollution, due to its
high solubility, and the emergence of resisting weeds. The negative impact of glyphosate residues
on the aquatic environment is a major concern due to the potential adverse effects to ecosystems
and human beings. Regarding to toxicity, there is a broadening controversy, but recent studies
have shown that glyphosate can affect phytoplankton and aquatic organisms. Vendrell et al. (2009)
observed a growth inhibition of 10% in green algae Scenedesmus subspicatus with 1.6 mg L−1
glyphosate acid treatments. Commercial formulations are often more toxic than glyphosate alone.
Tsui and Chu (2003) found a widespread commercial formulation to have seven folds higher
toxicity than glyphosate alone for green algae Selenastrum capricornutum. Pérez et al. (2007)
found that glyphosate can affect phytoplankton and periphyton community composition. Poletta
et al. (2009) revealed adverse effects of a commercial formulation of glyphosate on Caiman
latirostris DNA. A similar work by Guilherme et al. (2012) showed that commercial glyphosate
can increase DNA damage in teleost fish Anguilla anguilla.
At the same time, increasing and routine use of glyphosate has led to the emergence of weed
species tolerant or resistant to this active ingredient (Vitta et al., 2004;Yanniccari et al., 2012). This
implies environmental and economic costs, besides productivity losses (Binimelis et al., 2009).
For preventing or delaying the growth of glyphosate resistant weeds, a typical recommended
management strategy is to use herbicide mixtures (Diggle et al., 2003; Stewart et al., 2011).
Glyphosate is combined with herbicides of alternative modes of action and with soil residual
activity like 2,4-dichlorophenoxyacetic acid, commonly named 2,4-D (Miesel et al., 2012: Soltani
et al., 2009). 2,4-D is a phenoxy herbicide widely used in many countries to control broad leaf
weeds and it can also be easily transferred to natural water due to its high solubility (Kundu
et al., 2005). Ateeq et al. (2002) found that 2,4-D was cytotoxic and genotoxic for catfish Clarias
batrachus, as it causes micronuclei and erythrocytes alterations. Braga da Fonseca et al. (2008)
found that 2,4-D affects brain and muscle acetylcholinesterase (AChE) activity and metabolic
99
100 M. Mariani et al.

parameters of piava freshwater fish (Leporinus optusidem) due to the stress generated by this
herbicide.
The environmental pollution caused by pesticides has led to the development of efficient
treatment technologies for its degradation or removal. Among the different approaches to pesti-
cide removal, advanced oxidation processes (AOPs) have been recognized as especially effective
methods for this type of pollutants (Burrows et al., 2002; Ikehata and Gamal El-Din, 2006).
There are many studies which deal with the degradation of these herbicides in water with AOPs.
Glyphosate was studied under the ferrioxalate system (Chen et al., 2007), using photocatalytic
degradation with TiO2 (Echavia et al., 2009; Munner and Boxall, 2008; Shifu and Yunzhang,
2007), UV/H2 O2 (Mannasero et al., 2010). 2,4-D removal was studied using the Fenton’s reaction
(Pignatello, 1992), using ultraviolet radiation (Chamarro et al., 1993), UV/O3 (Prado et al., 1994),
electrochemical oxidation (Brillas et al., 2000; Oturan et al., 2000), the O3 /H2 O2 process (Yu
et al., 2006), the combined TiO2 /UVA/O3 and Fe(II)/UVA/O3 systems (Giri et al., 2008; Piera
et al., 2000), UV/H2 O2 (Alfano et al., 2001). All of these works aimed at the degradation of these
herbicides under the best operating conditions but considering just one contaminant at a time.
In aquatic environments, pesticides usually appear as complex mixtures instead of single con-
taminants (Relyea et al., 2009). For this reason, in order to approach to real pollution problems,
it is more important to test the mixture of these compounds rather than the individual chemical
separately (Chelme Ayala et al., 2010; Chioma Affam et al., 2012; Lapertot et al., 2007; Oller
et al., 2007).
The UV/H2 O2 method has some advantages regarding equipment and process, as hydrogen
peroxide is a relatively cheap chemical and the operative process is very simple. The aim of this
work was to evaluate the technical feasibility of the UV/H2 O2 process for the treatment of a mixture
of two herbicides: glyphosate and 2,4-D at the same concentration level. The influence of the
experimental conditions, such as initial hydrogen peroxide dose, initial pH, initial concentration
of the mixture of glyphosate and 2,4-D and incident radiation on the oxidation process, have been
examined. Toxicity assays (employing Vibrio fischeri) were also performed to monitor samples
at different stages of the treatment.

6.2 MATERIALS AND METHODS

6.2.1 Chemicals
The following reagents were used: (i) glyphosate (AccuStandard) as standard chromatographic,
(ii) glyphosate 95% provided by Red Surcos, (iii) 2,4 dichlorophenoxyacetic acid (Sigma-Aldrich)
as standard chromatographic, (iv) hydrogen peroxide (Ciccarelli p.a., >99%), (v) glycolic acid
(Merck), (vi) acetic acid (Anedra), (vii) 2,4-dichlorophenol (Aldrich), (viii) oxalic acid (Sigma-
Aldrich), (ix) potassium chloride (Merck), (x) sodium phosphate (Carlo Erba), and (xi) catalase
from bovine liver, >2000 units/mg (Fluka, 1 unit decomposes 1 mmol H2 O2 per minute at pH
7.0 and 25◦ C). Distilled water was used in all experiments.

6.2.2 Experimental setups and procedures


The degradation of the mixture of glyphosate and 2,4-D was carried out in a cylindrical reactor
made of Teflon® , with two parallel, flat windows made of quartz (VReactor = 110 cm3 ). The radi-
ation was produced with two Heraeus UVC lamps operated with an input power of 40 W each
(output power equal to 16 W = 3.4 × 10−5 Einstein s−1 ). They are low pressure mercury vapor
lamps (Germicidal type) with one single significant emission wavelength at 253.7 nm. Each win-
dow permitted the interposing of one shutter to block the passage of light when necessary. The
distance between the quartz windows and the lamps is 12 cm.
The photoreactor was operated as part of a recycling system that includes: a peristaltic pump,
a heat exchanger (for temperature control) and a well stirred tank with provisions for sampling
Degradation of a mixture of glyphosate and 2,4-D in water solution 101

Figure 6.1. Equipment setup (Zalazar et al., 2007).

Table 6.1. Experimental conditions.

Variable Value

Glyphosate initial concentration 15–70 mg L−1


2,4-D initial concentration 15–70 mg L−1
H2 O2 initial concentration 0–800 mg L−1
Incident radiation
(Einstein cm−2 s−1 ) × 109
Heraeus 40 W 22.4
Heraeus 40 W (with filter 43%) 9.8
Heraeus 40 W (with filter 18%) 4.1
Reaction time 6h
Temperature 20◦ C
Initial pH 3.5-7-10

and temperature measurements. The volume totalized a capacity of 1000 cm3 (VTank ). Figure 6.1
shows a schematic representation of the experimental setup.
The experiments were carried out changing the following variables: (i) initial glyphosate and
2,4-D concentrations in the mixture, (ii) initial hydrogen peroxide concentrations, (iii) initial
pH and (iv) incident radiation on the reactor windows measured with potassium ferrioxalate
actinometry (Murov et al., 1993) and calculated according to Zalazar et al. (2005) (Table 6.1).

6.2.3 Analytical measurements


Glyphosate was analyzed by ion chromatography with a suppressed conductivity detector and
employing an Ion Pac AG2A-SC guard column, an AS2A-SC separating column, and an ion
self-regenerating suppressor (Dionex). A solution of Na2 CO3 (7.2 mM) and NaOH (3.2 mM) was
used as eluent at a flow-rate of 0.6 mL min−1 . The injection volume was 20 µl. Prior to the liquid
chromatograph injection, catalase solution was added to each sample.
102 M. Mariani et al.

2,4-D and 2,4-DCP (2,4-dichlorophenol) were analyzed by high performance liquid chromatog-
raphy (Waters 1525 Binary HPLC Pump) with UV detector (Waters 2489 UV/Visible Detector),
using an analytical column of C18 reverse phase (SUPELCOSILTM LC-18 – 4.6 × 250 mm).
Acetonitrile (49%), Water (50%) and Acetic Acid (1%) were used as the mobile phase and the
separation was performed at a flow rate of 0.75 mL min−1 and 30◦ C. The fixed-wavelength UV
absorbance detector was set at 290 nm. The injection volume was 20 µl.
Oxalic and glycolic acids and final products (chloride and phosphate ions) were monitored
by ion chromatography using a solution of Na2 CO3 (1.3 mM) and NaHCO3 (1.4 mM) as eluent.
Acetic acid and glyoxylic acid were monitored by ion chromatography using a solution of NaB4 O7
(0.66 mM) as eluent.
pH was measured with a HI 98127 Hanna pH-meter (accuracy: ±0.1). Hydrogen peroxide was
analyzed using a colorimetric method following Allen et al. (1952) technique and employing a
Cary 100 Bio UV visible spectrophotometer. TOC was analyzed in order to compare glyphosate
and 2,4-D degradation rates with the total mineralization rate. The instrument used was a Shi-
madzu TOC-5000A. COD measurements were obtained using a reactor digestion method USEPA
approved (Hach Method 8000), within the range 0–150 mg L−1 COD.

6.2.4 Toxicity assay


Microtox Acute Toxicity Test was performed to evaluate the toxicity of the initial mixture of
glyphosate and 2,4-D and during the AOP assays. A Model 500 Analyzer (Strategic Diagnostic
Inc.) according to ASTM Standard Method D 5660-96 (2004) was used to measure the light
emitted by a specially selected freeze-dried strain of the marine bacterium Vibrio fischeri as a
result of its normal metabolic processes. Hydrogen peroxide present in the samples was removed
prior to every toxicity analysis using catalase after adjusting the sample pH between 6 and 8.
Microtox 81.9% screening test protocol was used for the toxicity assessment of samples.

6.2.5 Operation
The experimental run was started after every variable of the operating conditions had reached
its steady-state and/or uniformity. The photoreactor was filled with the working solution and
recirculation was established. At the same time that the lamp shutters where taken off the sample
a t = 0 was withdrawn. Afterwards, 25 mL samples were taken at different time intervals. After
finishing each run the photoreactor was carefully washed three times with tap water and twice
with deionized water.
A normal run lasted for 6 h. It must be noted that due to the type of equipment used in this
work (a recycle with a tank) this time does not represent the one effectively corresponding to the
irradiation time of the total system volume. Thus, the actual exposure to radiation must take into
account the ratio given by the photoreactor volume over the total volume (VReactor /VTotal = 0.11)
(Zalazar et al., 2007). This consideration should be taken into account when analyzing the reaction
times that are mentioned in the abscissas of several of the figures in this work.

6.3 RESULTS AND DISCUSSION

6.3.1 Preliminary runs


Two types of runs were made to investigate the effect of UV and H2 O2 separately. One group of
0
runs were made employing CGlyphosate = 29.5 mg L−1 and C2,4-D
0
= 29.1 mg L−1 without hydrogen
peroxide and using the 40 W Heraeus UV lamps during 6 h of reaction time (Fig. 6.2).
For glyphosate, minimal changes in concentrations were observed after 180 min and then
the concentration values remain constant. This is in agreement with the absorption spectrum of
glyphosate, at least in the range from 200 to 400 nm (Rueppel et al., 1977). For 2,4-D, if no H2 O2
was added in the solution a 33% decay of 2,4-D was observed after 6 hours of treatment.
Degradation of a mixture of glyphosate and 2,4-D in water solution 103

Figure 6.2. Concentration of glyphosate and 2,4-D in the mixture vs. time for a run without H2 O2 .

A similar set of runs were carried out with the following concentrations of glyphosate, 2,4-D
and H2 O2 : 26.6, 27.5 and 172.2 mg L−1 respectively and without UV radiation. No signals of
degradation were observed.

6.3.2 Effect of initial pH values


The pH can be an important factor for the reaction. The pH value, which permits to obtain the
highest degradation rate of a pollutant, depends on several conditions.
(i)The rate of photolysis of aqueous hydrogen peroxide increases when more alkaline conditions
are used. It has been suggested that HO−2 species might be responsible for the enhanced generation
of hydroxyl radicals under such conditions (pKa of H2 O2 = 11.6) since peroxide anion has a higher
molar napierian absorption coefficient at 253.7 nm (2.4 × 105 mol−1 cm2 ) than hydrogen peroxide
(1.86 × 104 mol−1 cm2 ). At the same time, a disproportion reaction of hydrogen peroxide also
occurs at alkaline pH values, as shown below (Ikehata and Gama El-Din, 2006):
H2 O2 + HO•2 −→ H2 O + O2 + HO• (6.1)

(ii) Lower operating pH (2.5–3.5) is usually preferred when radical scavengers such as carbon-
ate and bicarbonate ions are present in water. The oxidation efficiency is drastically reduced by
increasing pH above 5, since under these conditions bicarbonate ion is expected to successfully
scavenge HO• . An increase of bicarbonate concentration first lowers the hydroxyl radical concen-
tration (Eq. 6.2) and then reduces the destruction rate of the pollutant. The increase of pH above
7 transforms the bicarbonate into carbonate ion (Eq. 6.3), which has an even higher reactivity
towards HO• (Tang 2004).
HO• + HCO− 2− • •− −
3 /CO3 → HCO3 /CO3 + OH (6.2)
HCO•3 ⇔ CO•−
3
+
+ H where pKa = 7.9 (6.3)

(iii) Finally, another important factor is the pKa of the involved compound (Gogate and Pandit,
2004). The reactivity of a particular pollutant towards the HO• radical could depend on its ionized
form, which it is determined by the pH of the solution.
In this work, runs were performed at initial concentrations of glyphosate and 2,4-D of 30 mg L−1
and at an initial concentration of hydrogen peroxide of 150 mg L−1 . Three initial pH values
104 M. Mariani et al.

Figure 6.3. 2,4-D and glyphosate degradation at initial pH 3.5; 7 and 10. C0Glyphosate = 30 mg L−1 ; C02,4-D =
30 mg L−1 ; C0H2 O2 = 150 mg L−1 .

Figure 6.4. Different forms of glyphosate as a function of pH.

were chosen: the natural pH resulting from the chosen concentrations, pH = 3.5, pH = 7.0 and
pH = 10.0, artificially obtained by adding NaOH solution. The results are shown in Figure 6.3.
2,4-D was completely degraded after 2 hours of reaction for the three initial pH studied. The
highest decay rates were observed at acidic pH.
On the other hand, the initial pH plays a more important role in the degradation of glyphosate. It
was found that the variation of the initial pH solution had significant influence on the glyphosate
oxidation. When the initial pH value varied from 3.5 to 10.0 the glyphosate removal efficiency
after 6 hours of treatment increased from 51% (initial pH = 3.5) and 56% (initial pH 7.0) to 85%
for initial pH = 10.0.
Glyphosate contains three chemical groups -phosphate, amino and carboxylic- all of which
can be protonated and desprotonated as a function of pH (Fig. 6.4) (Haag and Yao, 1992; Munner
and Boxall, 2008).
Figure 6.4 shows that with a pH between 5.5 and 10.6 glyphosate has a net charge equal to
−2 because it has the three hydroxyl groups ionized and the amino group protonated (phosphate
−2, carboxylate −1, amino +1). The phosphonate group is totally ionized and, probably in this
ionic form, the breakdown of the C-P bond could be more favorable. Manassero et al. (2010)
proposed a plausible degradation path employing the H2 O2 /UVC process. In this proposal, the
initial glyphosate degradation begins with the breakdown of the C-P bond. In Section 6.3.7 of this
work “Formation of by-products and intermediates”, the degradation path is better explained.
Degradation of a mixture of glyphosate and 2,4-D in water solution 105

Figure 6.5. Effect of initial pH on the degradation of the mixture of 2,4-D and glyphosate. C0Glyphosate =
30 mg L−1 ; C02,4-D = 30 mg L−1 ; C0H2 O2 = 150 mg L−1 .

Figure 6.6. Effect of initial pH on the mineralization of TOC. C0Glyphosate = 30 mg L−1 ; C02,4-D = 30 mg L−1 ;
C0H2 O2 = 150 mg L−1 .

Figure 6.5 shows that initial pH = 10 is the best value to obtain high degradation levels of 2,4-D
and glyphosate mixture when using the UV/H2 O2 process.
It is also interesting to analyze the evolution of TOC at different initial pH because it provides
information about the overall kinetics of decomposition of the mixture of herbicides. Figure 6.6
shows how the initial pH affects the extent of the TOC mineralization. It can be seen that the rate
of mineralization is slightly better at an initial pH = 10. It is important to notice that the value of
TOC includes all the intermediate species formed during the oxidation process.
106 M. Mariani et al.

Figure 6.7. Glyphosate conversion percentage for 6 h reaction time vs. molar ratio and 2,4-D conversion
percentage for 1 h reaction time vs. molar ratio.

6.3.3 Effects of initial hydrogen peroxide concentration


There is usually an optimum concentration beyond which the presence of hydrogen peroxide is
detrimental to the degradation reaction due to a well-established scavenging action. The optimum
H2 O2 concentration depends on the amount and type of the pollutants in the water, i.e. on the
kinetic constant for the reaction between the free radicals and the pollutant and the rate constant
for the recombination reaction:
H2 O2 + HO• → HO•2 + H2 O (6.4)

The optimum concentration of the hydrogen peroxide may be established using laboratory
studies for the pollutant in question unless data are available in the literature with similar operating
conditions (Gogate and Pandit, 2004).
In addition to this, the use of high concentrations of hydrogen peroxide can affect the process due
to competing reactions. In Equation (6.4) hydroxyl radicals are consumed to form hydroperoxyl
radicals. This reaction produces an inhibitory effect for the degradation of contaminants because
HO•2 radical has lower oxidation capability. Also, HO• radicals can recombine according to:
HO•2 + HO• → H2 O + O2 (6.5)

Reactions (6.4) and (6.5) consumed hydroxyl radicals decreasing the oxidation probability of
the contaminant (Litter, 2005).
In this study the results for a fixed reaction time were analyzed in terms of glyphosate and 2,4-D
conversion percentages in the mixture vs. the initial molar ratio defined as r = C0H2 O2 /CTOC
0
, where
0
CTOC is the total organic carbon initial concentration corresponding to the mixture of pollutants.
Figure 6.7 shows that there is a range of optimum hydrogen peroxide concentrations. Both
glyphosate and 2,4-D showed the highest conversions for initial molar ratios between 4 and 10
(C0H2 O2 = 164–400 mg L−1 ).
The decay rate was optimized as the ratio increased to 10. When the ratio was further increased
up to 20, the degradation rate decreased.
For a run under the best experimental conditions for degradation, Figure 6.8 shows the temporal
progression for the concentrations of the participating species. The conversion of glyphosate was
almost 80% after 6 h. For 2,4-D the complete degradation was achieved at 2 h.
Note that for glyphosate the decay rate was significantly lower than for 2,4-D which indicates
that 2,4-D is easily degraded during the first stages of the reaction. Second-order reaction rate
Degradation of a mixture of glyphosate and 2,4-D in water solution 107

Figure 6.8. Concentrations vs. time. Experimental conditions: C0Glyphosate = 26.1 mg L−1 ; C02,4-D =
30.9 mg L−1 ; C0H2 O2 = 164.5 mg L−1 , initial pH = 10.

Figure 6.9. 2,4-D and glyphosate structures.

constants with HO• in aqueous solution (5 × 109 –1010 M−1 s−1 ) are observed for compounds
exhibiting (i) aromatic rings and/or carbon-carbon double bonds with electron-donating sub-
stituents, and /or (ii) aliphatic groups from which an H-atom can be easily abstracted. Significantly
slower rates are found only for compounds that do not exhibit any aromatic ring or carbon-carbon
double bond, and for aliphatic compounds with not easily abstractable H-atoms. Such H-atoms
include those that are bound to carbon atoms carrying one or several electronegative heteroatoms
or groups. Thus the differences in the reactivity between glyphosate and 2,4-D could be explained
by their structure differences (Fig. 6.9). The second order rate constants values are 5 × 109 M−1 s−1
and 1.8 × 108 M−1 s−1 for 2,4-D and glyphosate respectively (Haag and Yao, 1992).

6.3.4 Effect of glyphosate and 2,4-D initial concentrations


Glyphosate and 2,4-D degradation for different initial concentrations are shown in Figure 6.10.
For the mixture of 15.9 mg L−1 of 2,4-D and 16.7 mg L−1 of glyphosate, removal percentages
were 100% for 2,4-D after 60 min of reaction time, and 99% for glyphosate after 6 hours of
treatment. For the mixture of 30.9 mg L−1 of 2,4-D and 26.1 mg L−1 of glyphosate, the percentages
of removal obtained were 92.2% for 2,4-D after 60 min, and 85.6% for glyphosate, after 6 hours.
And for the mixture of 60.1 mg L−1 of 2,4-D and 68.3 mg L−1 of glyphosate with the same reaction
times, the percentages obtained were 73.5% and 32.6% respectively.
For a better understanding of the influence of the initial concentration on the herbicide mixture
degradation, it is possible to represent the effect of initial concentration on the degradation initial
108 M. Mariani et al.

Figure 6.10. Effect of glyphosate and 2,4-D initial concentrations: concentration vs. reaction time.
(a) 2,4-D; (b) Glyphosate.

Figure 6.11. Effect of glyphosate and 2,4-D initial concentration on the initial degradation rate.

rate. As shown in Figure 6.11, the initial rate of glyphosate increases as the pesticide concentration
is raised from 15 to 30 mg L−1 , but it goes through a maximum and then decreases when the initial
glyphosate concentration is higher than 30 mg L−1 . On the other hand, the initial rate of 2,4-D
always increases in the range of 0–60 mg L−1 of initial concentrations.
When working with the mixture, the degradation rate corresponding to glyphosate is lower at
the beginning of the reaction and, afterwards, the decay rate increases. The degradation of 2,4-D
does not present this type of behavior. To confirm this result, additional experiments were made at
different ratios of initial concentrations of 2,4-D and glyphosate but maintaining the same initial
molar ratio r. Figure 6.12a shows that there is not too much difference in the degradation rates
of 2,4-D when it is alone or when it is mixed with glyphosate. In contrast, Figure 6.12b shows the
degradation rates for glyphosate alone and for two different mixtures of both herbicides. In these
experiments, the behavior mentioned was confirmed, since when the concentration of 2,4-D is
low the degradation rate of glyphosate increases more quickly.

6.3.5 Effect of variations in the incident UV spectral fluence rate


at the irradiated reactor walls
These results provide a precise, indirect way to know the effect produced by the power of the
employed radiation sources on the degradation rates of each of the studied compounds. For a
Degradation of a mixture of glyphosate and 2,4-D in water solution 109

Figure 6.12. Dimensionless concentrations as a function of time for: (a) (•): 2,4-D alone: C02,4-D =
29.5 mg L−1 ; C0H2 O2 = 159.6 mg L−1 ; initial pH = 10; r = 4 (): Mixture of glyphosate + 2,4-
D: C0Glyphosate = 26.1 mg L−1 ; C02,4-D = 30.9 mg L−1 ; C0H2 O2 = 164.5 mg L−1 ; initial pH = 10;
r = 4 (b) (•): Glyphosate alone: C0Glyphosate = 29.1 mg L−1 ; C0H2 O2 = 76.4 mg L−1 ; initial
pH = 10; r = 4 (): Mixture of glyphosate + 2,4-D: C0Glyphosate = 39.1 mg L−1 ; C02,4-D =
14.6 mg L−1 ; C0H2 O2 = 153.9 mg L−1 ; initial pH = 10; r = 4 (): Mixture of glyphosate + 2,4-
D: C0Glyphosate = 26.1 mg L−1 ; C02,4-D = 30.9 mg L−1 ; C0H2 O2 = 164.5 mg L−1 ; initial pH = 10;
r = 4.

Figure 6.13. Effect of variations in the incident UV spectral fluence rate at the irradiated reactor walls.
(a) Dimensionless 2,4-D concentration vs. time (b) Dimensionless glyphosate concentra-
tion vs. time. The spectral fluence rate at the irradiated reactor walls is the parameter.
C0Glyphosate = 30 mg L−1 ; C02,4-D = 30 mg L−1 ; C0H2 O2 = 150 mg L−1 .

photochemical reaction, this information is very important and can define if, from the point of
view of the use of the applied energy, the process may or may not be feasible.
The time concentration evolution during the decomposition of each of the components of the
mixture of 2,4-D and glyphosate applying the UV/H2 O2 process with the same lamps, but using
neutral density filters that permitted the passage of 43 and 18% of the spectral UV fluence rate
at the reactor wall are shown separately in Figure 6.13a,b, respectively.
For 2,4-D the conversion percentages reached for a reaction time of 2 h, employing lamps with
an input power of 40 W were 100% while for the same lamps with filters (43 and 18%) were 64
and 57% respectively (Fig. 6.13a). Figure 6.13b shows that for a reaction time of 6 h, using the
110 M. Mariani et al.

Figure 6.14. Evolution of TOC removal ratio of the mixture of herbicides in aqueous solution during the
UV/H2 O2 process C0Glyphosate = 30 mg L−1 ; C02,4-D = 30 mg L−1 ; initial pH = 10.

40 W lamp a glyphosate conversion of 85% was reached, while conversions for the lamps with
filters (43 and 18%) were 46 and 35%, respectively.
It can be seen that in this case, the observed differences are very significant, permitting to
conclude that the employed radiation power is a very significant variable for both reactions.

6.3.6 Total organic carbon (TOC) evolution


It is very important to know the extent of the mineralization of the mixture of herbicides during
and at the end of the UV/H2 O2 process. Mineralization was followed by measuring the TOC. The
TOC removal ratio was calculated according to:
C0TOC − CtTOC
TOC removal reaction = (6.6)
C0TOC

with C0TOC is the total organic carbon at t = 0 and CTOC


t
is the total organic carbon during the
reaction time.
The results are shown in Figure 6.14. The highest percentage of TOC removal (∼70%) was
achieved for initial molar ratios between 4 and 10 (C0H2 O2 = 164–400 mg L−1 ).

6.3.7 Formation of by-products and intermediates


In order to evaluate the oxidation of the mixture of glyphosate and 2,4-D and to obtain a better
understanding of the reaction mechanism involved, a by-product evaluation is needed. However,
given the complex variety of the reaction products that could be generated in this process, an
exhaustive identification and quantification of all the intermediate species is very difficult. For
this reason, this study was focused on the major stable by-products of the reaction.
For 2,4-D degradation, 2,4-DCP was the first intermediate identified as an important stable
oxidation intermediate, which indicates that the degradation of 2,4-D begins from the rupture
of the alcoxy bond (Yu et al., 2006). Further reactions of 2,4-DCP with loss of chloride ions
form carboxylic acids which are transformed into oxalic acid. It is important to notice that when
2,4-DCP is completely degraded (approximately 120 min of reaction), chloride ions concentra-
tions reached asymptotic values. This would indicate that after 2 hours of reaction there are not
chlorinated intermediates.
Degradation of a mixture of glyphosate and 2,4-D in water solution 111

Figure 6.15. Formation of by-products in 2,4-D degradation (a) 2,4-DCP and chloride ion (b) oxalic,
acetic, glyoxylic and glycolic acids. C0Glyphosate = 26.1 mg L−1 ; C02,4-D = 30.9 mg L−1 ;
C0H2 O2 = 164.5 mg L−1 ; initial pH = 10.

Figure 6.16. Evolution of chloride in the mixture of glyphosate and 2,4-D degradation
C0Glyphosate = 26.1 mg L−1 ; C02,4-D = 30.9 mg L−1 ; C0H2 O2 = 164.5 mg L−1 ; initial pH = 10.

Aliphatic carboxylic acids have generally been reported to be the final organic degradation
products, prior to mineralization of 2,4-D, during the application of AOPs. Ion-exclusion chro-
matography revealed the presence of glycolic acid, glyoxylic acid, oxalic acid and acetic acid in
the solution (Brillas et al., 2000; Oturan et al., 2000). The results are shown in Figures 6.15a
and 6.15b.
Finally, 2,4-D is mineralized to yield carbon dioxide and chloride ions as final products.
Figure 6.16 shows that for each mol of 2,4-D that is decomposed, two mol of chloride ions
appears at each reaction time.
Several by-products have been proposed for the oxidation of glyphosate with different AOPs
like AMPA, sarcosine and glycine (Chen et al., 2007; Munner and Boxall, 2008). In Mannassero
et al. (2010) only glycine was first identified and, later on, nitrate and phosphate were found as
final by-products.
112 M. Mariani et al.

Figure 6.17. Evolution of phosphate in the mixture of glyphosate and 2,4-D degradation
C0Glyphosate = 26.1 mg L−1 ; C02,4-D = 30.9 mg L−1 ; C0H2 O2 = 164.5 mg L−1 ; initial pH = 10.

In our experiments, sarcosine and AMPA were not identified. Only glycine was quantified but
its concentration was very low.
The attack of HO• radical to glyphosate leads to the formation of a carbon centered radical

CH2 -NH+ −
2 -CH2 -COO , and the C-P cleavage leads to formation of phosphate. Thus, it could
be said that degradation of glyphosate may be observed and it may be followed by the phosphate
evolution. Figure 6.17 shows that for each mol of glyphosate that is decomposed, one mol of
phosphate appears at each reaction time.

6.3.8 Toxicity and chemical oxygen demand assays


When employing AOPs for safe removal of pesticides, the efficiency of the process should be
evaluated by the degree of detoxification. It is known that some intermediates generated during
photooxidation could be even more toxic and/or recalcitrant than the parent compound, thus
toxicity measurements are an essential final task. This information is very important for the
disposal of the treated effluent into the environment or into a conventional bio treatment plant.
Ecotoxicity bioassays provide important information to detect the degree of chronic effects of
substances on representative organisms and were designed to evaluate variations in toxicity. Bac-
terial assays are often used because they are relatively quick and simple an also because bacteria
are an integral part of the ecosystem. The microbial test most frequently used for monitoring
water quality is the Microtox test, which uses the bioluminescence of the marine bacterium Vibrio
fischeri (Parvez et al., 2006).
The luminescence of V. fischeri is directly associated with respiratory activity and this provide
a good indicator of metabolic actions and, therefore, of the general toxicity of a compound or of
a mixture of compounds (Ince et al., 1999).
In experiments with Vibrio fischeri where glyphosate was used, Bonnet et al. (2007) obtained
an EC 50,30min = 21.25 mg L−1 after 30 min of incubation while Hernando et al. (2007) found that
glyphosate was harmful with an EC 50,30min = 44.2 mg L−1 . On the other hand, Zona et al. (2003)
reported that 2,4-D is a moderately toxic compound with an EC 50,30min = 59 mg L−1 for V. fischeri.
The toxicity was calculated by using the Microtox data acquisition software as a percentage of
inhibition of luminescence upon contact times of the samples at 5 and 15 min. The results for a
typical run are shown in Figure 6.18.
Degradation of a mixture of glyphosate and 2,4-D in water solution 113

Figure 6.18. Toxicity (percentage of inhibition of Vibrio fischeri) and TOC of samples treated by the
UV/H2 O2 process. C0Glyphosate = 27.9 mg L−1 ; C02,4-D = 29.4 mg L−1 ; C0H2 O2 = 146.9 mg L−1 ;
initial pH = 10.

At the beginning of the treatment, the percentage of bacteria inhibition was high (∼85%
inhibition) and after 10 min the toxicity is still elevated. This might be due to the formation of
2,4-DCP, which has been described as toxic and more resistant to oxidative degradation than
2,4-D. EC 50 for 2,4-DCP was approximately 5 mg L−1 and it was determined by the Microtox
bioassay (Drzewicz et al., 2004). Thereafter, the toxicity decreased quickly, but between 60 min
and 120 min of treatment a slight rise in toxicity can be observed. It could be attributed to the
accumulation of chlorinated oxidation intermediates, which can be even more toxic, and the high
concentrations of glyphosate remain in the solution. After 120 min, when total dechlorination has
been achieved, the toxicity decreases gradually, following the TOC changes. At the end of the
process, the mixture showed almost not relevant toxicity (∼10% inhibition) and nearly 76% of
TOC removal, confirming the capability of the UV/H2 O2 process to detoxify the herbicide mixture.
From these results we can conclude that it is not necessary to complete the mineralization of the
mixture of glyphosate and 2,4-D because intermediate organics compounds present after 420 min
of treatment have very low toxicity.
Chemical oxygen demand (COD) is widely used as an assay to identify the characteristics of
wastewater. For the optimal operating conditions found in this work, a result of 87% of COD
removal was obtained after 6 h of treatment. COD was also measured to evaluate the average
oxidation stage (AOS) of the mixture of glyphosate and 2,4-D. In complex solutions consisting of
original compounds and their by-products, AOS can be used to estimate the oxidative ability and
also provides indirect information of the biodegradability of the treated solution (Jimenez et al.,
2011). AOS takes values between +4 for CO2 , the most oxidized state of C, and −4 for CH4 , the
most reduced state of C.
Then, the value of AOS was calculated according to:
4 × (TOC-COD)
AOS = (6.7)
TOC
where TOC and COD are expressed in molar concentration.
The results for COD measurements and the calculated AOS are presented in Figure 6.19. Also,
values of TOC are shown. AOS increased sharply at the beginning of the treatment (from −0.5 to 2)
and after approximately 60 min remained almost constant until the end of the treatment. These
results show that at the beginning more oxidized organic intermediates are formed and after a
certain time a plateau is reached, indicating that the oxidative nature of the organic intermediates
114 M. Mariani et al.

Figure 6.19. TOC, COD and AOS during degradation of the herbicide mixture C0Glyphosate = 27.9 mg L−1 ;
C02,4-D = 29.4 mg L−1 ; C0H2 O2 = 146.9 mg L−1 ; initial pH = 10.

formed did not varied substantially. TOC measurements during the process are correlated with
AOS. At the beginning there is a decrease in TOC due to degradation of 2,4-D mainly. This stage
corresponds to the increase in the AOS; i.e., when total dechlorination has been achieved. Then,
AOS remains almost stable due to the formation of carboxylic acids that have slow mineralization
rates.
In any event, it is important to remark that a significant reduction of COD and TOC of the
herbicide mixture solution was achieved at the end of the treatment.

6.4 CONCLUSIONS

This experimental study aims at researching the feasibility of the UV/H2 O2 process for degrading
a mixture of glyphosate and 2,4-D. The following conclusions could be drawn from the present
study:
• The best conditions found in this experimental work to treat an aqueous solution containing
typical concentrations of a mixture of herbicides (30 mg L−1 of glyphosate and 30 mg L−1 of
2,4-D ) are: hydrogen peroxide concentrations C0H2 O2 = 164–400 mg L−1 , an initial pH = 10
and germicidal lamps with an incident radiation of 22.4 × 109 Einstein cm−2 s−1 .
• A complete degradation of 2,4-D was achieved after 2 hours of treatment. Glyphosate was
degraded up to 85% when the reaction was extended for 4 additional hours. This degradation
was accompanied with removal values corresponding to TOC = 76% and COD = 87%.
• The Microtox assays performed indicated that the toxicity of the mixture of glyphosate and
2,4-D was significantly reduced after the treatment. Moreover, it is not necessary to reach a
complete mineralization in order to obtain almost non-toxic end products.

ACKNOWLEDGEMENTS

The authors would like to thank to Universidad Nacional del Litoral, Consejo Nacional de Inves-
tigaciones Científicas (CONICET) and Agencia Nacional de Promoción Científica y Tecnológica
for their financial support. Also thanks to Margarita Herman for the English language editing.
The technical assistance of Eng. Claudia Romani is gratefully appreciated.
Degradation of a mixture of glyphosate and 2,4-D in water solution 115

REFERENCES

Alfano, O., Brandi, R. & Cassano, A.: Degradation kinetics of 2,4-D in water employing hydrogen peroxide
and UV radiation. Chem. Eng. J. 82 (2001), pp. 209–218.
Allen, A., Hochanadel, C., Ghormley, J. & Davis, T.: Decomposition of water and aqueous solutions under
mixed fast neutron and gamma radiation. J. Phys. Chem. 56 (1952), pp. 575–586.
ASTM Standard D5660-96, Standard test method for assessing the microbial detoxification of chemi-
cally contaminated water and soil using a toxicity test with a luminescent marine bacterium. ASTM
International, West Conshohocken, PA, doi:10.1520/D5660-96R04, http://www.astm.org, 2004.
Ateeq, B., Abul Farah, M., Niamat Ali, M. & Ahmad, W.: Induction of micronuclei and erythrocyte alterations
in the catfish Clarias bratachus by 2,4 dichlorophenoxyacetic acid and butachlor. Mutation Res. 518
(2002), pp. 135–144.
Binimelis, R., Pengue, W. & Monterroso, I.: “Transgenic treadmill”: Responses to the emergence and spread
of glyphosate-resistant johnsongrass in Argentina. Geoforum 40 (2009), pp. 623–633.
Braga da Fonseca, M., Glusczak, L., Moraes, B., Cavalheiro de Menezes, C., Pretto, A., Tierno, M.A.,
Zanella, R., Ferreira Goncalves, F. & Loro, V.: The 2,4-D herbicide effects on acetylcholiesterase activity
and metabolic parameters of piava freshwater fish (Leporinus obtusidens). Ecotoxicol. Environ. Saf. 69
(2008), pp. 416–420.
Brillas, E., Calpe, J.C. & Casado, J.: Mineralization of 2,4-D by advanced electrochemical oxidation
processes. Water Res. 34:8 (2000), pp. 2253–2262.
Bonnet, J.L., Bonnemoy, F., Dusser, M. & Bohatier, J.: Assessment of the potential toxicity of herbicides and
their degradation products to nontarget cells using two microorganisms, the bacteria Vibrio fischeri and
the ciliate Tetrahymena pyriformis. Environ. Toxicol. 22:1 (2007), pp. 78–91.
Burrows, H.D., Canle, M.L., Santaballa, J.A. & Steenken, S.: Reaction pathways and mechanisms of
photodegradation of pesticides. J. Photochem. Photobiol. B 67 (2002), pp. 71–108.
Chamarro, E. & Esplugas, S.: Photodecomposition of 2,4-Dichlorophenoxyacetic acid: Influence of pH.
J. Chem. Technol. Biotechnol. 57 (1993), pp. 273–279.
Chelme-Ayala, P., Gamal El-Din, M. & Smith, D.: Kinetics and mechanism of the degradation of two
pesticides in aqueous solutions by ozonation. Chemosphere 78 (2010), pp. 557–562.
Chen, Y., Wu, F., Lin, Y., Deng, N., Bazhin, N. & Glebov, E.: Photodegradation of glyphosate in the
ferrioxalate system. J. Hazard. Mater. 148 (2007), pp. 360–365.
Chioma Affam, A., Kutty, S.R. & Chaudhuri, M.: Solar photo-fenton induced degradation of combined
chlorpyrifos, cypermethrin and chlorothalonil Pesticides in aqueous solution. I. J. Chem. Environ. Eng. 6
(2012), pp. 167–173.
Diggle, A.J., Neve, P.B. & Smith, F.P.: Herbicides used in combination can reduce the probability of herbicide
resistance in finite weed populations. Weed Res. 43 (2003), pp. 371–382.
Drzewicz, P., Nalecz-Jawecki, G., Gryz, M., Sawicki, J., Bojanowska-Czajka, A., Głuszewski,W., Kulisa,
K., Wołkowicz, S. & Trojanowicz, M.: Monitoring of toxicity during degradation of selected pesticides
using ionizing radiation. Chemosphere 57 (2004), pp. 135–145.
Echavia, G.R.M., Matzusawa, F. & Negishi, N.: Photocatalytic degradation of organophosphate and
phosphonoglycine pesticides using TiO2 immobilized on silica gel. Chemosphere 76 (2009), pp. 595–600.
FAO, FAOESTAT. Internet site: http://faostat.fao.org/site/567/default.aspx#ancor (accessed September
2012).
Giri, R.R., Ozaki, H., Takanami, R. & Taniguchi, S.: A novel use of TiO2 fiber for photocatalytic ozonation
of 2,4-dichlorophenoxyaceticacid in aqueous solution. J. Envirom. Sci. 20 (2008), pp. 1138–1145.
Gogate, P. & Pandit, A.: A review of imperatives technologies for wastewater treatment II: Hybrid methods.
Adv. Environ. Res. 8 (2004), pp. 553–597.
Guilherme, S., Gaivao, I., Santos, M.A. & Pacheco, M.: DNA damage in fish (Anguilla anguilla) exposed to
a glyphosate-based herbicide.Elucidation of organ-specificity and the role of oxidative stress. Mut. Res.
Genet. Toxicol. Environ.Mutagen. 743 (2012), pp. 1–9.
Haag, W.R. & Yao, C.C.: Rate constants for reaction of hydroxyl radicals with several drinking water
contaminants. Environ. Sci. Technol. 26 (1992), pp. 1005–1013.
Hach Method 8000 Oxygen Demand Chemical, for water, wastewater and seawater. Internet site:
http://www.hach.com/epa (access 20 August 2012).
Hernando, M.D., De Vettori, S., Martínez Bueno, M.J. & Fernández-Alba, A.R.: Toxicity evaluation with
Vibrio fischeri test of organic chemicals used in aquaculture. Chemosphere 68 (2007), pp. 724–730.
Ikehata, K. & Gamal El-Din, M.: Aqueous pesticide degradation by hydrogen peroxide/ultraviolet irradiation
and Fenton-type advanced oxidation processes: A review. J. Environ. Eng. Sci. 5 (2006), pp. 81–135.
116 M. Mariani et al.

Ince, N.H., Dirilgen, N., Apikyan, I. G., Tescanli, G. & Ustun, B.: Assessment of toxic interactions of heavy
metals in binary mixtures: A statistical approach. Arch. Environ. Contam. Toxicol. 36 (1999), pp. 365–372.
Jiménez, M., Oller, I., Maldonado, M.I., Malato, S., Hernández-Ramìrez, A., Zapata, A. & Peralta-
Hernández, J.M.: Solar photo-Fenton degradation of herbicides partially dissolved in water. Catal. Today
161 (2011), pp. 214–220.
Kundu, S., Pal, A. & Dikshit, A.K.: UV induced degradation of herbicide 2,4-D: Kinetics, mechanism and
effect of various conditions on the degradation. Sep. Purif. Technol. 44 (2005), pp. 121–129.
Lapertot, M., Ebrahimi, S., Dazio, S., Rubinelli, A. & Pulgarin, C.: Photo-Fenton and biological integrated
process for degradation of a mixture of pesticides. J. Photochem. Photobiol. A 186 (2007), pp. 34–40.
Litter, M.: Tecnologías avanzadas de oxidación: Tecnologías solares. In: M. Blesa & J. Blanco Gálvez
(eds): Tenologías solares para la desinfección y descontaminación del agua. Proyect FP6-510603. Sixth
Framework Program of of the European Union, Argentina, 2005, pp. 183–194.
Manassero, A., Passalia, C., Negro, A., Cassano, A. & Zalazar, C.: Glyphosate degradation in water
employing the H2 O2 /UVC process. Water Res. 44 (2010), pp. 3875–3882.
Miesel, J.R., Renz, M.J., Doll, J.E. & Jackson, R.D.: Effectiveness of weed management methods in estab-
lishment of switchgrass and a native species mixture for biofuels in Wisconsin. Biomass Bionergy 36
(2012), pp. 121–131.
Muneer, M. & Boxall, C.: Photocatalyzed degradation of a pesticide derivative glyphosate in aqueous
suspensions of titanium dioxide. Int. J. Photoenergy (2008), Article ID 197346.
Murov, S., Carmichael, I. & Hug, G.: Handbook of photochemistry. Second ed., Marcel Dekker, New York,
1993.
Oller, I., Malato, S., Sánchez-Pérez, J.A., Maldonado, M.I. & Gassó, R.: Detoxification of wastewater
containing five common pesticides by solar AOPs–biological coupled system. Catal. Today 129 (2007),
pp. 69–78.
Oturan, M.A.: An ecologically e€ective water treatment technique using electrochemically generated
hydroxyl radicals for in situ destruction of organic pollutants: Application to herbicide 2,4-D. J. Appl.
Electrochem. 30 (2000), pp. 475–482.
Parvez, S, Venkataraman, C. & Mukherji, S.: A review on advantages of implementing luminescence
inhibition test (Vibrio fischeri) for acute toxicity prediction of chemicals. Environ. Int. 32 (2006), pp.
265–268.
Perez, G.L., Torremorell, A., Mugni, H., Rodríguez, P., Solange Vera, M., do Nascimento, M., Allende, L.,
Bustingorry, J., Escaray, R., Ferrara, M., Izaguirre, I., Pizarro, H., Bonetto, C., Morris, D.P. & Zagarese,
H.: Effects of herbicide roundup on freshwater microbial communities: A mesocosm study. Ecol. Appl.
17 (2007), pp. 2310–2322.
Piera, E., Calpe, J.C., Brillas, E., Doménech, X. & Peral, J.: 2,4-Dichlorophenoxyacetic acid degradation by
catalyzed ozonation: TiO2 /UVA/O3 and Fe(II)/UVA/O3 systems. Appl. Catal. B 27 (2000), pp. 169–177.
Pignatello, J.: Dark and photoassisted Fe3+ -catalyzed degradation of chlorophenoxy herbicides by hydrogen
peroxide. Environ. Sci. Technol. 26 (1992), pp. 944–951.
Poletta, G.L, Larrieta, A., Kleinsorge, E. & Mudry, M.E.: Genotoxicity of the herbicide formulations
Roundup (glyphosate) in broad-snouted caiman (Caiman latirostris) evidenced by the comey assay and
the micronucleus test. Mutation Res., Genet. Toxicol. Environ.Mutagen. 672 (2009), pp. 95–102.
Prado, J., Arantegui, J., Chamarro, E. & Esplugas, S.: Degradation of 2,4-D by ozone and light. Ozone-Sci.
Eng. 16 (1994), pp. 235–245.
Relyea, R.: A cocktail of contaminants: How mixtures of pesticides at low concentrations affect auatic
commnuties. Oecologia 159 (2009), pp. 363–376.
Rueppel, M.L., Brightwell, B.B., Schaefer, J. & Marvel, J.: Metabolism and degradation of glyphosate in
soil and water. J. Agric. Food Chem. 25 (1977), pp. 517–528.
Shifu, C. & Yunzhang, L.: Study on the photocatalytic degradation of glyphosate by TiO2 photocatalyst.
Chemosphere 67 (2007), pp. 1010–1017.
Soltani, N., Shropshire, C. & Sikkema, P.: Sensitivity of winter wheat to preplant and preemergence
glyphosate tankmixes. Crop Prot. 28 (2009), pp. 449–452.
Stewart, C.L., Nurse, R.E., Van Eerd, L.L., Vyn, R.J. & Sikkema, P.H.: Weed control, environmental impact,
and economics of weed management strategies in glyphosate-resistant soybean. Weed Technol. 25 (4)
(2011), pp. 535–541.
Tang, W.Z.: Physicochemical treatment of hazardous wastes. Lewis Publishers, USA, Florida, 2004.
Tsui, M.T. & Chu, L.M.: Aquatic toxicity of glyphosate-based formulations: Comparison between different
organisms and the effects of environmental factors. Chemosphere 52 (2003), pp. 1189–1197.
Degradation of a mixture of glyphosate and 2,4-D in water solution 117

Vendrell, E., Ferraz, D. G., Sabater, C. & Carrasco, J. M.: Effect of glyphosate on growth of four fresh-
water species of phytoplankton: A microplate bioassay. Bull. Environm. Contam. Toxicol. 82 (2009),
pp. 538–542.
Vitta, J.I., Tuesca, D. & Puricelli, E.: Widespread use of glyphosate tolerant soybean and weed community
richness in Argentina. Agri. Ecosyst. Environ. 103 (2004), pp. 621–624.
Yanniccari, M., Istilart, C., Giménez, D.O. & Castro, A.M.: Glyphosate resistance in perennial ryegrass
(Lolium perenne L.) from Argentina. Crop Protection 32 (2012), pp. 12–16.
Yu, Ying-hui, Ma, Jun & Hou, Yan-jun.: Degradation of 2,4-dichlorophenoxyacetic acid in water by ozone-
hydrogen peroxide process. J. Environ. Sci.18:6 (2006), pp. 1043–1049.
Zalazar, C.S., Labas, M.D., Martín, C.A., Brandi, R.J., Alfano, O.M. & Cassano, A.E.: The extended use of
actinometry in the interpretation of photochemical reaction engineering data. Chem. Eng. J. 109 (2005),
pp. 67–81.
Zalazar, C., Labas, M., Brandi, R. & Cassano. A.: Dichloroacetic acid degradation employing hydrogen
peroxide and UV radiation. Chemosphere 66 (2007), pp. 808–815.
Zona, R. & Solar, S.: Oxidation of 2,4-dichlorophenoxyacetic acid by ionizing radiation: Degradation,
detoxification and mineralization. Radiat. Phys. Chem. 66 (2003), pp. 137–143.
This page intentionally left blank
CHAPTER 7

Degradation of perchlorate dissolved in water by a combined


application of ion exchange resin and zerovalent iron nanoparticles

Luis Cumbal, Daniel Delgado & Erika Murgueitio

7.1 INTRODUCTION

Perchlorate (ClO− 4 ) is an anion that has a chloride atom with an oxidation state of positive seven.
This compound holds a tetrahedral geometry, formed with four oxygen atoms, and its negative
charge is stabilized by mesomeric effects (Urbansky, 2002). Due to this symmetrical configura-
tion, perchlorate anion shows a high chemical stability. Also, ClO− 4 exhibits high water solubility
and low vapor pressure, properties that converts to perchlorate in an important contaminant of
water systems (ITRC, 2005; Urbansky, 2002). Perchlorate is produced either by natural or anthro-
pogenic sources. A place where ClO− 4 naturally occurred is at saltpeter deposits of Chile (Dadoly,
2005). In contrast, the anthropogenic perchlorate contamination comes from the use of perchlorate
salts and perchloric acid; mainly in the manufacturing of fertilizers, dyes and paints, explosives,
fireworks, etc. (Urbansky, 2002). Salts of perchlorate are particularly used by the ammunition
and military industry or industry related to aerospace programs. For example, ammonium per-
chlorate is utilized as a solid oxidant in fuel of missiles, ammunitions and in propulsion systems
of rockets (Rogers, 1998; Urbansky, 2002). The intensive use of perchlorate compounds builds up
contamination mostly on surface and in ground bodies of water; developing large plumes which
can reach up to 14.5 km of length. For this reason, this toxic compound is considered culprit of
awful environmental impacts (EPA, 2007).
Perchlorate can cause several affections on human health. When this chemical is orally ingested,
it can trigger a decrease on the thyroid hormones production, because this compound is able to
inhibit the transfer of iodine from blood to thyroid glandule. A deficit of thyroid hormones for
a long time can produce hypothyroidism and changes on the metabolism of elder people. In
infants and neonates, a deficiency of thyroid hormones, even during a short term, is very risky,
because these hormones take part in the development of several endocrine organs, such as the
brain (EPA-IRIS, 2005; ATSDR, 2008). Bearing in mind these affections, humans can orally take
0.7 micrograms of perchlorate per kilogram of body weight on daily basis (0.7 µg kg−1 d−1 ) and
the maximum allowed concentration in drinking water is 24.5 µg L−1 (EPA, 2007).
Moreover, perchlorate does not chemically react with other substances thus the process of
removing this compound from surface and ground water is exceptionally difficult and expensive
(EPA, 2012). Currently, there are some treatments used to remove perchlorate from water. Among
these treatments, ion exchange, adsorption with activated carbon, composting, permeable reactive
barriers, phytoremediation, electrodialysis, microfiltration, and reverse osmosis have been used
with a modest or any success (EPA-FFRRO, 2005).
Scientific literature reports that zerovalent iron nanoparticles (nZVI) have been successfully
used on the elimination or reduction of chlorinated solvents and inorganic ligands (He and Zhao,
2005). The transformation of halogenated organic compounds (HOCs) using zerovalent iron is
considered one of the most innovative technologies in environmental remediation. In addition,
nZVI covered with palladium have been also used to eliminate HOCs. These nanomaterials are
capable of removing chloride atoms from some aliphatic chlorinated compounds and mixtures
of polychlorinated biphenyls (PCBs) through reductive reactions (Wang and Zhang, 1997). Con-
versely, researchers have demonstrated that polymeric ion exchange resins selectively remove
119
120 L. Cumbal, D. Delgado & E. Murgueitio

perchlorate from contaminated waters. In this process, perchlorate replaces anions of less affinity
in the resin. However, this method does not destroy perchlorate but only remove it from water
(Dozier et al., 2004). In this context, this study is aimed to develop an innovative technique which
combines the benefits of ion exchange for uptaking the toxic compound from water and the zero
valent nanoparticles to degrade perchlorate to chloride.

7.2 EXPERIMENTAL SECTION

7.2.1 Chemicals
For testing, chemicals of ACS grade were used: anion exchange resin (Purolite A530E); cation
exchange resin (Purolite C104); ferrous sulfate heptahydrate (Aldrich, 99%); sodium borohydride
(Sigma Aldrich, 98%); carboxymethylcellulose (CMC, MW = 90,000, Aldrich); sodium chloride
(Fisher Scientific, 100% ACS); absolute ethanol (Panreac chemicals, 99.5%); sodium perchlorate
(Aldrich, >98%) sodium hydroxide (Merck, ≥99%); hydrochloric acid (Fisher Scientific, 37.3%
certified ACS plus); oxalic acid (Aldrich, ≥99%); sodium sulfate (Sigma Aldrich, 98%); sodium
bicarbonate (Sigma Aldrich, 98%).

7.2.2 Procedures
7.2.2.1 Preparation of nanoparticles
Stabilized nZVI were prepared following the protocol developed by He and Zhao (2005) and
modified in the Centro de Investigaciones Científicas de la Escuela Politécnica del Ejercito
(ESPE). The method used CMC as stabilizer. For the synthesis, it was prepared a CMC–Fe(II)
solution containing 0.8% (w/w) CMC and Fe(II) at a desired concentration (1.0, 1.5, and 2.0 g L−1
as Fe). NaBH4 solution was stoichiometrically added at a flow rate of 10 mL min−1 to reduce
Fe2+ to Fe0 . The preparation was carried out in a 500 mL flask connected to a vacuum line; a
non-magnetic stirring was kept during the addition of the reducing agent. All aqueous solutions
were prepared with deionized (DI) water previously purged with purified nitrogen gas (N2 ) for
15–30 min to remove dissolved oxygen.
When the gas (hydrogen) evolution ceased (after approximately 15 min) the reduction of Fe2+
to Fe0 ended up and the nanoparticles were ready to be used as metallic agent. For comparison,
nonstabilized Fe nanoparticles were prepared following similar procedures but in the absence of
carboxymethylcellulose.

7.2.2.2 Physical characterization of nanoparticles


Transmission electrode microphotographs were obtained by using a FEI Tecnai Spirit Twin trans-
mission electron microscope (TEM) operated at 80 kV. The following general procedures were
followed for the TEM measurements. First, a 40 µL of aqueous sample containing nZVI was
placed on a Formvar-carbon coated copper 400 mesh grid. The droplet was allowed for full con-
tact/spreading on the grid. The droplet was then wicked away with a filter paper to remove the
excess volume of sample, and the grid was allowed to air-dry for 4 minutes. Second, the residual
nanoparticles attached to the grid were then imaged and photographed using the TEM. The TEM
image was analyzed using a specialty image processing software named TIA 1994–2010 (TEM
Image and Analysis) to obtain particle morphology and distribution.
The mean hydrodynamic size of the prepared nZVI was obtained with a dynamic light scattering
(DLS) submicrometer particle sizer (HORIBA LB-550, Irvine, CA) at a measurement angle of
90◦ (Laserdiode, wavelength of 650 nm, 15 mW). The DLS data was processed with the software
package HORIBA LB-550 to yield the volume weighted size distribution. In all measurements, at
least duplicate samples (∼1 mL each) of 0.1 g L−1 of Fe were analyzed. Freshly prepared samples
of 1.0, 1.5, and 2.0 g L−1 Fe were diluted to 0.1 g L−1 before analysis.
Degradation of perchlorate dissolved in water 121

Table 7.1. Salient properties of the A530E ion exchanger resin.

Property Description/value

Type Strong base


Matrix Macroporous styrene cross-linked with divinylbenzene
Functional group Quaternary ammonium
Ionic form as shipped Cl− form
Total capacity 0.6 eq L−1 (Cl− form)
Moisture retention 49–55% (Cl− form)
Particle size range 300–1200 µm
Specific gravity 1.4 g mL−1
Manufacturer Purolite

7.2.2.3 Conditioning of ion exchange resins and loading with perchlorate


A Purolite A530E resin was used to remove perchlorate from water samples (Table 7.1). The
conditioning was carried out through consecutive washes with 1 mol L−1 NaOH solution, DI
water and 1.0 mol L−1 HCl and a final wash with 1.0 mol L−1 NaCl. Five bed volumes of the
washing solutions were used. The conditioned resin was loaded with a concentrated solution of
perchlorate (40% w/w) during 24 h at 180 rpm.

7.2.2.4 Kinetic tests


For kinetic studies, tests were performed contacting 100 mg of wet resin loaded with perchlorate
with 80 mL of 1.0 g L−1 of Fe0 nanoparticles stabilized with 0.8% CMC and with 10% NaCl
and without sodium chloride during 12 h at 25 and 95◦ C. Every 2 hours a 2.0 mL liquid sample
from each test was collected to measure the amount of soluble perchlorate. The degradation of
perchlorate was stopped adding 2.0 mL of 3.0 mol L−1 HCl solution into the collected samples. The
aqueous samples containing perchlorate and Fe0 nanoparticles were pretreated with concentrated
6.0 mol L−1 NaOH to precipitate out iron atoms and filtered with 0.2 µm PVDF filters. The
residual of dissolved iron in the filtrate was removed with a Purolite C104 resin. The remaining
liquid phase was filtered through 0.45 µm PVDF filters before injecting into a chromatograph.

7.2.2.5 Degradation of perchlorate


Tests were performed contacting 100 mg of wet resin loaded with only perchlorate with 40 mL
of 1.0, 1.5, and 2.0 g L−1 of nZVI during eight hours. Two temperatures were tested for the
experiments (25 and 95◦ C). When the treatment finished, 3.0 mL of aqueous samples from each
test were collected to measure the amount of soluble perchlorate as on the kinetic tests. The amount
of perchlorate that was not reduced and still remained in the polymeric resin was measured as
follows: (i) polymeric beads were separated from the solution of Fe particles by filtration and
regenerated with 10% NaCl + 30% CH2 CH3 OH for 1 hour. Spent regenerant samples were filtered
and diluted with DI water before injecting into the ionic chromatograph. The following tests were
conducted to assess the degradation of perchlorate: (i) with 10% NaCl + 10% CH3 CH2 OH, (ii)
with only 10% NaCl, (iii) with only 10% ethanol, (iv) without any regenerant and (v) control
samples.

Degradation tests of perchlorate using waters containing sulfate (SO2− 4 ), bicarbonate (HCO3 ),
and oxalic acid (C2 H2 O4 ) (an organic matter surrogate) were conducted under otherwise the same
experimental conditions as above. Three groups of water samples were prepared. The first group
was prepared with 200 mg L−1 ClO− 4 , 120 mg L
−1
SO2−
4 , and 100 mg L
−1
HCO− 3 , the second
group with 200 mg L−1 ClO− 4 and 30 mg L −1
C H O
2 2 4 , and the third group with 200 mg L−1
ClO− 4 , 120 mg L
−1
SO2−
4 , 100 mg L
−1
HCO− 3 and 30 mg L
−1
C2 H 2 O 4 .
122 L. Cumbal, D. Delgado & E. Murgueitio

7.2.3 Chemical analysis


Chemical analysis of perchlorate was conducted in a DIONEX ICS-1100 chromatograph,
equipped with a 4 mm ASRS-300 self-regenerating suppressor, a 4 mm AG16 pre-column, a
4 mm AS16 analytical column and a 50 µL loop. As eluent, a 35 mM NaOH solution was used.
Working conditions during the analytical procedure were: a suppression current of 100 mA and a
flow rate of 1.0 mL min−1 . Under these experimental conditions, the detection limit for perchlorate
was 60 to 80 µg L−1 .

7.3 RESULTS AND DISCUSSION

7.3.1 Physical characterization of nanoparticles


Figure 7.1 compares the transmission electron microphotographs of nZVI prepared in presence
(A) and absence (B) of 0.8% (w/w) CMC stabilizer. Figure 7.1B shows that in the absence of the
stabilizer the resulting nanoparticles do not appear as discrete nanoscale particles but form bulkier
aggregates (flocs) with varying density. The size of the denser flocs can be well greater than 1 µm.
This type of aggregation was associated with the magnetic forces among the Fe particles (Zhang
and Manthiram, 1997). In contrast, the CMC-stabilized Fe particles shown in Fig. 7.1A appeared
to be clearly discrete and well dispersed. A similar TEM image of discrete particles was observed
by other researchers for nonmagnetic nanoparticles (Zhang and Manthiram, 1997). Evidently, the
presence of CMC avoided agglomeration of the resulting nanoparticles and therefore maintained
the high surface area and reactivity of the particles.
Figure 7.2 compares the mean particle size of 1.0, 1.5, and 2.0 g L−1 of Fe nanoparticles pre-
pared with 0.8% (w/w) CMC and obtained with a DLS particle analyzer. Figure 7.2 shows that
at 1.0 g L−1 of Fe, the mean size is smaller (8.52 ± 3.02 nm) compared with the other concen-
trations. On the contrary, at 2.0 g L−1 of Fe, the mean size is greater (12.95 ± 5.53 nm). Similar,
mean particle sizes were reported by Cook (2009). He found sizes of nZVI that are typically in
the range of 5 to 40 nm thus securing a high reactivity. For discrete nanoparticles such as those
shown in Figure 7.1A, the surface area (S) is inversely proportional to particle radius (r) through
the Equation (7.1) (Kecskes et al., 2003):
r = 3(ρS)−1 (7.1)

where ρ is the density of Fe (7870 kg m−3 ). Based on mean radii of 4.26, 5.58, and 6.48 nm
(for 1.0, 1.5, and 2.0 g L−1 of Fe0 ), the surface areas for the CMC-stabilized iron particles were

Figure 7.1. TEM images of 1 g L−1 of nZVI: (A) prepared with a CMC and under inert conditions; the
scale bar represents 100 nm, (B) prepared without CMC but under inert conditions; the scale
bar represents 100 nm.
Degradation of perchlorate dissolved in water 123

calculated to be ∼90, 68, and 59 m2 g−1 . It is obvious that as particle size decreases its specific
surface area increases as well as its reactivity.

7.3.2 Kinetic tests


Figure 7.3 compares the transient degradation of perchlorate with and without NaCl at 25 and
95◦ C. Figure 7.3 shows that perchlorate achieves the highest degradation after seven to eight
hours of treatment. At 25◦ C, the degradation proceeds slowly but when the temperature is raised
to 95 ◦ C the reaction is accelerated. It is noteworthy to see that during the first 2 or 4 hours
of tests performed with 10% NaCl, perchlorate concentration rapidly increases in the solution
phase. Cl− ions that are in excess in solution displace the toxic anion from the resin. Almost
similar trends show tests conducted without NaCl; nevertheless, desorption of perchlorate from
the resin is much slower due to the initial absence of Cl− ions in solution. As ZVI nanoparticles
diffuse through the resin, contact and reduce ClO−4 ions bound on the ion-exchanger to chloride;

Figure 7.2. Sizes of nZVI prepared with 0.8% CMC (MW = 90,000, T = ∼20◦ C).

Figure 7.3. Transient degradation of soluble perchlorate using 1.0 g L−1 of nZVI stabilized with 0.8% CMC
with and without NaCl at 25 and 95◦ C.
124 L. Cumbal, D. Delgado & E. Murgueitio

however, these ions are slowly produced. As the concentration of chloride is build up in the ZVI
nanoparticles solution, the perchlorate desorption is even further promoted.
In the perchlorate-laden solution, ZVI nanoparticles undergo the following competitive
reactions (Cao et al., 2005; Gu et al., 2003):
4Fe0 + 2H2 O → Fe2+ + H2 + 2OH− (7.2)

ClO− 0 + −
4 + 4Fe + 8H → Cl + 4Fe
2+
+ 4H2 O (7.3)
As the reactions proceed, the surface of the ZVI nanoparticles is progressively oxidized to iron
oxides (Xiong et al., 2007). In the experimental pH range (7.3–7.7), the resultant iron oxides can
rapidly adsorb ClO− 4 from water, resulting in the observed drop in perchlorate concentration in the
solution phase such that its concentration gradient is reversed until it reaches steady state. Studies
on destruction of perchlorate present in water and ion-exchange brine using ZVI nanoparticles
reveal sorption/desorption process governed the kinetic reaction of perchlorate (Xiong et al.,
2007).
Overall, better desorption and degradation of perchlorate is achieved with 10% NaCl at 95◦ C.
At higher temperature, the porous network of the resin expands and easily releases perchlorate
(DeSilva and Koebel, 2000). Unbound perchlorate is solubilized and underwent the same process
explained above. Previous studies reported that at higher temperatures, the activation energy
needed to degrade perchlorate is easily satisfied (Cao et al., 2005; Oh, 2010; Oh et al., 2006a;
2006b). Another researcher reported that a good degradation of perchlorate was achieved after
seven hours of treatment with 1.8 g L−1 of Fe0 nanoparticles stabilized with 1.2% CMC using
6% (w/w) NaCl, in a range of temperatures between 25 and 110◦ C (Xiong, et al., 2007). A
similar kinetic profiles were observed by Moore et al. (2003) and Moore and Young (2005) using
commercial Fe nanoparticles.

7.3.3 Degradation of perchlorate


Figures 7.4 and 7.5 show perchlorate degradation in the presence and the absence of regenerant
solutions (NaCl or CH3 CH2 OH). Figure 7.5 shows that degradation of perchlorate is greater in
all experiments conducted at 95◦ C. At high temperatures, there is an increase of perchlorate
desorption from the resin and a rapid reduction of the toxic compound with nZVI, as discussed
above. Perchlorate degradation using 10% NaCl and 10% ethanol is limited because the toxic
anion is not easily released from the resin. On the contrary, with a solution containing 10%
ethanol and 10% NaCl, perchlorate shows a better degradation. The addition of the alcohol
enhances desorption of perchlorate because it decreases the dielectric constant of the regenerant
solution. Gu et al. (2001) previously reported that a regenerant solution containing 0.35 mol L−1
FeCl3 , 2 mol L−1 HCl, and 35% ethanol was effective in the elution of ClO− 4 compared with only
a 2 mol L−1 NaCl solution. In this study 10% ethanol yields good results on the regeneration of
the resin, while concentrations higher than 10% produce an early oxidation of Fe nanoparticles,
compromising the reduction of perchlorate to chloride.
Perchlorate degradation is expected to take place due to the large specific surface areas and the
excellent reactivity of nZVI. Fresh nZVI used on tests had 90, 68, and 59 m2 g−1 for 1.0, 1.5, and
2.0 g L−1 of Fe0 , respectively. Consequently, the resulting large surface areas assure a high chemi-
cal reactivity. It has been reported that as surface areas of nZVI increase their chemical, reactivity
increases as well (He and Zhao, 2005; Tiwari et al., 2008). However, as seen in Figures 7.4 and
7.5, perchlorate degradation is also increased as the concentration of Fe nanoparticles increases.
This is because nanoparticles population is proportional to Fe concentration. At 2.0 g L−1 of Fe0 ,
the number of nanoparticles is almost doubled compared to at 1.0 g L−1 of Fe0 . Accordingly, the
surface area increases favoring degradation of more toxic compound.
The degradation of perchlorate to chloride due to the redox power of nZVI follows a series
of successive reactions as described by Equations (7.3) and (7.4). Nevertheless, this degradation
reaction is not simple and it goes through a number chemical intermediate compounds such as
Degradation of perchlorate dissolved in water 125

Figure 7.4. Perchlorate degradation using different concentrations of nZVI after 8 hours of treatment at
25◦ C.

Figure 7.5. Perchlorate degradation using different concentrations of nZVI after 8 hours of treatment at
95◦ C.

chlorate, chlorite, hypochlorite, and/or chloride (Eq. 7.4) (Cao et al., 2005; Gu et al., 2003;
Rikken et al., 1996).
ClO4 → ClO− − −
3 → ClO2 → ClO → Cl

(7.4)

7.3.4 Effect of competing ions and organic matter on the degradation of perchlorate
Tests with water containing perchlorate and competing ions demonstrated that A530E resin still
uptakes perchlorate selectively. Nevertheless, there is a decrease of a 4% on the perchlorate uptake
compared to those water samples containing only the pollutant (Fig. 7.6). Scientific literature
reports that affinity of an ion exchanger varies with solubility, size of the hydrated radius, and
charge of ions to be removed. In case of A530E resin, the polystyrene matrix cross-linked with
polyvinyl benzene chains is hydrophobic thus the hydration barrier is reduced. As a result, a high
126 L. Cumbal, D. Delgado & E. Murgueitio

Figure 7.6. Uptake of perchlorate from contaminated water by the A530E resin.

interaction between hydrophobic ions and the methyl functional groups of the resin takes place
(Xiong, 2007).
It is well known that perchlorate, a monovalent anion, has a relatively small hydration energy
(−259 kJ mol−1 ) compared to the majority of anions present in natural waters and, as a result,
perchlorate ion is more hydrophobic. In contrast, sulfate and bicarbonate have higher hydration
energies (−1103 kJ mol−1 and −388 kJ mol−1 , respectively) (Brown et. al., 2000). Consequently,

SO2−4 and HCO3 are hydrophilic and their affinity for the resin is limited. High selectivity
of A530E resin for perchlorate ions is then attributed to the structure of the functional groups
(Purolite A 530E). The reactive sites of the resin are formed with three methyl groups. Thus, a
localized hydrophobic environment is developed affecting sorption of hydrophilic anions (SO2− 4
and HCO− 3 ) (Gu et al., 2001). Besides, the quaternary ammonium groups generate a high charge
separation with the multivalent species. Accordingly, sorption of multivalent anions is scarce. Due
to these properties, the A530E resin exhibits high specificity for anions of a large size, with plain
charge, and relatively low hydration energy as shown by perchlorate and pertechnetate (Gu and

Coates, 2006; Gu et al., 2001). In fact, the presence of SO2− 4 and HCO3 in water contaminated
with perchlorate, does not interfere with the capacity of the A530E resin to uptake perchlorate,
neither with the reduction of this pollutant when the resin gets in contact with the suspension
of nZVI and the regenerants. Under these conditions, perchlorate is degraded up to 92.5% on
average (Fig. 7.7).
Results of tests using water with 30 mg L−1 oxalic acid (organic matter surrogate) and
200 mg L−1 perchlorate show that the presence of the organic compound indeed affected the
overall treatment. The uptake of perchlorate by the resin was 19% less compared to results with
only perchlorate (Fig. 7.6). This loss of selectivity is attributed to the presence of the oxalic acid
in water samples. Large molecules or big ions released from the organic compounds (uncharged
oxalic acid and oxalate anion from the ionization of the oxalic acid) that diffuse through the
polymeric matrix of the resin might be retained by two mechanisms: (i) ionic bonding between
organic anions and the positively charged sites on the resin and (ii) hydrophobic bonding between
organic matter and the polymeric matrix due to van der Waals attractive forces (Purolite, 1999).
Also, the presence of organic matter in the contaminated water not only interferes with the per-
chlorate uptake capacity of the resin but also with the contaminant degradation when the resin
gets in contact with the suspension of Fe nanoparticles and the regenerants. In fact, the total
degradation of perchlorate was 6.4% less compared with those water samples without organic
Degradation of perchlorate dissolved in water 127

Figure 7.7. Perchlorate degradation using 2.0 g L−1 of Fe0 nanoparticles after eight hours of treatment at
95◦ C.

matter (Fig. 7.7). The organic matter that binds to the resin and hinders the release perchlorate
from the reactive groups triggers this interference (Purolite, 1999). In addition, the presence of
oxygen atoms in the oxalic acid molecule might contribute to the rapid oxidation of nZVI.
Findings of tests using water with perchlorate, competing anions, and organic matter, all
together, show an uptake of perchlorate around 77% (Fig. 7.6). This is principally attributed to
the presence of organic matter in the water samples. Perchlorate degradation in these tests is on
average 85.8%, namely 9.7% lower than its counterparts tested with the ion exchange resin loaded
with only perchlorate (Fig. 7.7). Similarly, the amount of chloride generated from the degradation
of perchlorate also decreased to 90% (data not shown).

7.4 CONCLUSIONS

The combined use of an ion exchange resin and nZVI fosters the removal of perchlorate from
contaminated water. Perchlorate captured by the A530E resin is successfully reduced to chlo-
ride through the application of Fe0 particles simultaneously with sodium chloride and ethanol
regenerant solution. It is believed that the degradation (reduction) of the contaminant is due to
the high chemical activity of nZVI. This novel remediation system that is based on ion exchange
and chemical reduction reactions can be readily used to recover surface and ground water bodies
contaminated with perchlorate with less time and economically. In fact, the degradation of per-
chlorate immobilized by the A530E resin, takes place in 8 hours when small volumes of nZVI are
applied. Furthermore, the regeneration capability of the resin facilitates its reuse in other cycles
of perchlorate sorption-desorption, making the process inexpensive.
The application of chemical regenerants (NaCl or CH3 CH2 OH) to desorb perchlorate from the
resin plays an important role on the contaminant degradation. Results of tests carried out with 10%
NaCl + 10% CH3 CH2 OH show high perchlorate degradation. It is obvious that both regenerants
speed up desorption of perchlorate from the resin and make possible the solubilization of the
contaminant in the suspension of nZVI. Accordingly, an excellent contact between perchlorate
and zerovalent iron nanoparticles is promoted. On the other hand, temperature is a critical factor
on the degradation of perchlorate. Results of tests conducted at 95◦ C show a substantial increase
on the degradation of the contaminant. High temperature in the suspension of the Fe0 particles
provides the activation energy needed to break down Cl-O bonds on the perchlorate.
The A530E resin selectively uptakes perchlorate even if competing ions are present in water
at high concentrations. The presence of 120 mg L−1 SO2− 4 and 100 mg L
−1
HCO− 3 in the loading
128 L. Cumbal, D. Delgado & E. Murgueitio

aqueous phase results in a decrease of only 4% on the perchlorate uptake. High selectivity of the
resin for perchlorate ions is associated with the structure of the functional groups. The reactive
sites of the resin are formed by three methyl groups thus a localized hydrophobic environment is
developed.
Regardless of A530E resin selectivity for perchlorate, the presence of organic matter (oxalic
acid: C2 H2 O4 ) indeed affects its uptake capacity. Big ions coming from the organic matter (oxalate
anion: C2 O2−4 ) are bound to the reactive groups of the resin through ion exchange interactions
while large organic matter molecules (uncharged oxalic acid) are attached to the polymeric matrix
of the resin byVan der Waals forces. As a result, 30 mg L−1 of oxalic acid turn out a 19% of decrease
on the perchlorate uptake by the resin. Nevertheless, the total perchlorate degradation when the
resin gets in contact with 2.0 g L−1 of nZVI and 10% NaCl + 10% C2 H5 OH at 95◦ C, decreases
only 5%. Organic matter retained on the reactive groups and on the matrix of the polymeric resin
complicates desorption and the solubilization of perchlorate. Besides, the presence of oxygen
atoms in the oxalic acid molecule might contribute to the rapid oxidation of nZVI.

ACKNOWLEDGEMENTS

Authors express their gratitude to Prof. Arthur Kney from Lafayette College for the donation of
laboratory chemicals and supplies and to Mr. Jeff Shoemaker for his support at the beginning of
the study. We also appreciate the help given by Dr. Alexis Debut with the TEM microphotographs.

REFERENCES

ATSDR (Agency for Toxic Substances & Disease Registry): ToxFAQs for perchlorates. 2008,
http://www.atsdr.cdc.gov (accessed December 2011).
Brown, G.M., Bonnesen, P.V., Moyer, B.A., Gu, B., Alexandratos, S.D., Patel, V. & Ober, R.: Efficient
treatment of perchlorate contaminated groundwater with bifunctional anion exchange resins. In: E.T.
Urbansky (ed): Perchlorate in the environment. Kluwer Academic/Plenum Publishers, New York, NY,
USA, 2000, pp.155–164.
Cao, J., Elliott, D. & Zhang, W.-X.: Perchlorate reduction by nanoscale iron particles. J. Nanopart. Res. 7
(2005), pp. 499–506.
Cook, S.M.: Assessing the use and application of zero-valent iron nanoparticle technology for remediation
at contaminated sites. Jackson State University, Jackson, MS, USA, 2009.
Dadoly, J.: Low concentrations of perchlorate detected again in surface water, groundwater. Pendleton, OR,
USA, 2005, http://www.deq.state.or.us/er/perchloratesites.htm (accessed November 2011).
DeSilva, F. & Koebel, B.: Temperature effects on ion exchangers. USA, 2000, http://www.resintech.com/pdf/
somelikeithot.pdf (accessed November 2011).
Dozier, C., Melton, H., Hare, F., Porter, O. & Lesikar, B.: Drinking water problems: Perchlorate. TX, USA,
http://texaswater.tamu.edu/resources/factsheets/l5468perchlorate.pdf (accessed November 2011).
EPA: Technical fact sheet–Perchlorate. 2012, http://www.epa.gov/fedfac/pdf/technical_fact_sheet_
perchlorate.pdf (accessed July 2012).
EPA, Federal Facilities Restoration and Reuse Office (EPA-FFRRO). Emerging contaminant – Perchlorate.
2007, www.epa.gov (accessed November 2011).
EPA, Integrated Risk Information System (EPA-IRIS). Perchlorate and perchlorate salts. 2005,
http://www.epa.gov/iris/subst/1007.htm (accessed November 2011).
EPA, Federal Facilities Restoration and Reuse Office (EPA-FFRRO). Perchlorate treatment technology
update. 2005, http://www.epa.gov/tio/download/remed/542-r-05-015.pdf (accessed November 2011).
Gu, B., Brown, G.M., Maya, L., Lance, M.J. & Moyer, B.A.: Regeneration of perchlorate loaded anion
exchange resins by a novel tetrachloroferrate displacement technique. Environ. Sci. Technol. 35:16 (2001),
pp. 3363–3368.
Gu, B. & Coates, J. (eds): Perchlorate environmental occurrence, interactions and treatment. Springer Inc,
New York, USA, 2006.
He, F. & Zhao, D.: Preparation and characterization of a new class of starch-stabilized bimetallic nanoparticles
for degradation of chlorinated hydrocarbons in water. Environ. Sci. Technol. 39:9 (2005), pp. 3314–3320.
Degradation of perchlorate dissolved in water 129

ITRC, Interstate Technology Regulatory Council: Perchlorate: Overview of issues, status, and remedial
options. USA, 2005, itrcweb.org/Guidance/GetDocument?documentID=60 (accessed November 2011).
Kecskes, L.J., Woodman, R.H., Trevino, S.F., Klotz, B.R., Hirsch, S.G. & Gersten, B.L.: Characterization of
a nanosized iron powder by comparative methods. KONA 21 (2003), pp. 143–150.
Moore, A.M. & Young, T.M.: Chloride interactions with iron surfaces: Implications for perchlorate and
nitrate remediation using permeable reactive barriers. J. Environ. Eng. 131:6 (2005), pp. 924–933.
Moore, A.M., De Leon, C.H. &Young, T.M.: Rate and extent of aqueous perchlorate removal by iron surfaces.
Environ. Sci. Technol. 37:14 (2003), pp. 3189–3198.
Oh, S.-Y.: Enhanced reduction of perchlorate by zero-valent iron: Effect of temperature, pH, and buffering
capacity. J. Geochem. Explor. 13:4 (2010), pp.119-126.
Oh, S.-Y., Chiu, P.C., Kim, B.J. & Cha, D.K.: Enhanced reduction of perchlorate by elemental iron at elevated
temperatures. J. Hazard. Mater. B 129 (2006a), pp. 304–307.
Oh, S.-Y., Cha, D.K., Chiu, P.C. & Kim, B.J.: Zero-valent iron treatment of RDX-containing and perchlorate-
containing wastewaters from an ammunition-manufacturing plant at elevated temperatures. Water Sci.
Technol. 54:10 (2006b), pp. 47–53.
Purolite: Purolite Technical Bulletin: Organic fouling of anion ion exchange resins. 1999, http://www.
purolite.com/relid/606346/isvars/default/customized/uploads/pdfs/organicfouling.pdf (accessed Septem-
ber 2011).
Purolite: Purolite®A530E. Macroporous strong base anion exchange resin. http://www.purolite.com/default.
aspx?RelID=606288&ProductID=333 (accessed August 2011).
Rikken, G.B., Kroon, A.G.M & van Ginkel C.G.: Transformation of (per)chlorate into chloride by a newly
isolated bacterium: Reduction and dismutation Appl. Microbiol. Biotechnol. 45 (1996), pp. 420-426.
Rogers, D.E.: Perchlorate user and production information. Department of the Air Force, Air Force Materie.
Command Law Office, Wright-Patterson Air Force Base, OH, USA, 1998.
Tiwari, D.K., Behari, J. & Sen, P.: Application of nanoparticles in wastewater treatment. World Appl. Sci. J.
3:3 (2008), pp. 417–433.
Urbansky, E.T.: Perchlorate as an environmental contaminant. Environ. Sci. Poll. Res. 9:3 (2002),
pp. 187–192.
Wang, Ch.-B. & Zhang, W.-X.: Synthesizing nanoscale iron particles for rapid and complete dechlorination
of TCE and PCBs. Environ. Sci. Technol. 31:7 (1997), pp. 2154–2156.
Xiong, Z.: Destruction of perchlorate and nitrate by stabilized zero valent iron nanoparticles and immobi-
lization of mercury by a new class of iron sulfide nanoparticles. PhD Thesis, Auburn University, Auburn,
AL, USA, 2007.
Xiong, Z. Zhao, D. & Pan, G.: Rapid and complete destruction of perchlorate in water and ion-exchange
brine using stabilized zero-valent iron nanoparticles. Water Res. 41:15 (2007), pp. 3497–3505.
Zhang, L. & Manthiram, A.: Chains composed of nanosize metal particles and identifying the factors driving
their formation. Appl. Phys. Lett. 70 (1997), pp. 2469–2471.
This page intentionally left blank
CHAPTER 8

Eco-friendly approach for Direct Blue 273 removal from


an aqueous medium

Pamela Yanina González Clar, Gustavo Levin, María Victoria Miranda &
Viviana Campo Dall’ Orto

8.1 INTRODUCTION

Synthetic dyes are complex aromatic compounds extensively used in the textile, leather, cosmet-
ics, food, paper and pulp industries. These compounds contain functional groups such as azo,
anthraquinone, triphenylmethane, heterocyclic or phthalocyanine as chromophore.
Dyes in wastewater cause serious environmental pollution. Some of them undergo chemical as
well as biological changes, consume dissolved oxygen, and destroy aquatic life. Others are very
stable to photo-degradation, temperature, oxidizing agents and microbial attack, making them
recalcitrant compounds. They are sometimes fused with heavy metals of toxic and inhibitory
nature (Venkata Mohan et al., 2005).
Synthetic dyes with azo bonds (–N=N–) represent the largest group of industrial colorants
produced annually, and are resistant to the biodegradation in aerobic conditions (Dos Santos et al.,
2007). On the other hand, the azo bond can be reduced by biotreatment in anaerobic conditions,
giving rise to aromatic amines which can be toxic and carcinogenic (Chung and Cerniglia, 1992).
Due to a general concern about the treatment of wastewater, a large number of projects have
been carried out to find more efficient methodologies for the treatment of contamination. Table
8.1 summarizes the recent achievements in the different technologies developed for the removal
of azo dyes using simulated effluents and real textile industry effluents.
The traditional approaches for wastewater treatment from dyeing industry include chemical
oxidation (Ruppert et al., 1994), coagulation/sedimentation, filtration (Majewska-Nowak et al.,
1989), membrane processes, and adsorption (Haque et al., 2011; Khaled et al., 2009; Ramakrishna
and Viraraghavan, 1997; Wawrzkiewicz, 2011). Each of these methods has its own advantages
and drawbacks which have been extensively reviewed (Cooper, 1995; Ferreira-Leitao et al.,
2007).
The advanced oxidation processes are based on the generation of highly oxidizing species,
such as the HO• radical, which has a high potential (E o approx. 2.80 V) and can promote the
degradation of a wide range of contaminants in minutes. Hydroxyl radicals can be generated
by reactions involving strong oxidizers, such as ozone, hydrogen peroxide and ultraviolet light,
Fenton reagent, ultraviolet light in presence of titanium dioxide and zinc oxide, ultrasound, or
electrochemical processes (Ahmed Basha et al., 2011; Baig and Liechti, 2001; Elias et al., 2011;
Mohajerani et al., 2009; Solis et al., 2012).
Bioresource technology is an acceptable option due to its ecofriendly nature, and production of
less sludge (Ferreira-Leitao et al., 2007; Park et al., 2007). Enzymes may transform pollutants to
diminish their toxicity, to increase water solubility allowing their further microbial degradation
or to promote insolubility and its subsequent removal from the industrial waste stream.
Peroxidases and polyphenol oxidases participate in the degradation of a broad range of sub-
strates even at very low concentrations, sometimes in presence of redox mediators, but large scale
exploitation has not been achieved due to the low stability of the enzymatic activity in biological
131
132 P.Y.G. Clar et al.

Table 8.1. Summary of methodologies for dye treatment.

Processes Concluding remarks

Advanced oxidation processes (Oller et al., 2011)


Fenton and photo-Fenton Successful treatment of textile dyes and commercial surfactants
(solar light or UV)
TiO2 photocatalysis (solar light High removal efficiency for reactive dyes
or UV) with or without H2 O2
Ozonation Formation of by-products which increased toxicity
of the formulation
TiO2 -assisted photocatalysis Complete discoloration and TOC reduction over 60%;
Ozonation 60% discoloration and negligible TOC reduction
Biosorbents (Srinivasan and Viraraghavan, 2010)
Fungi Biodegradation is the mechanism for living cells, and
biosorption for living and dead cells. High degree of
discoloration in minutes – days
Bacteria Extracellular solute biosorption by dead/inactive cells.
The chemical functional groups of the cell wall play vital roles
in this mechanism. Biodegradation processes may be
anaerobic, aerobic or involve a combination of the two.
High degree of discoloration in minutes – days
Chitosan It has a polycationic structure and high removal capacities for
anionic dyes such as acid, reactive and direct dyes
Algae Relatively high surface area and high binding affinity due to
cell wall properties. Other mechanisms are bioconversion
and biocoagulation due to long chain biopolymers
Enzymatic degradation (Solis et al., 2012)
Anaerobic microbial degradation There is a cleavage of the azo linkage with the aid of an
of azo dyes anaerobic azoreductase, and electron transfer by a redox
mediator that acts as an electron shuttle between the
extracellular dye and the intracellular reductase
Oxidative degradation of azo Laccases have potential in bioremediation due to their
dyes catalyzed by peroxidases non-specific oxidation capacity, lack of a requirement for
and phenoloxidases cofactors and use of readily available oxygen as an
electron acceptor.
Peroxidases use H2 O2 as substrate
Advanced oxidation processes (Solis et al., 2012)
combined with microbiological
processes
Ozonation + microbiological It improves sludge settleability and wastewater
treatment biodegradability and achieves good color removal of dyes
because it attacks conjugated azo bonds even at low
concentrations
Sonication + microbiological or Sonolysis degrades the dye into smaller compounds, thereby
enzymatic treatment enhancing biodegradability in shorter treatment
Fenton oxidation + This combination requires a 4- to 5-fold lower amount of
microbiological treatment Fenton’s reagent, reduces the process time, discolors
high concentration of dye, mineralizes and detoxifies the effluents.
When used in combination with microbiological methods,
Fenton is effective only as pre-treatment
Electrochemical oxidation + Good results in the degradation of azo dyes, with high
microbiological processes discoloration and COD removal
UV/H2 O2 + microbiological The combination reduces the cost of electricity consumption
processes.
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 133

Scheme 8.1. Chemical structure of DB273.

materials and high cost of purification (Jamal et al., 2011; Torres et al., 2003; Venkata Mohan
et al., 2005).
The partial degradation of different azo-dyes was achieved with manganese peroxidase, a
heme-containing glycoprotein that requires Mn2+ as cofactor and hydrogen peroxide as an oxidant
(Husain, 2010; Xiaobin et al., 2007). The mechanism of degradation involves both symmetrical
and asymmetrical azo-bond cleavages, and leads to products of lower molecular weight.
In this work we studied the discoloration of aqueous solutions of Direct Blue 273 (DB273), an
azo-dye usually employed in paper industry whose discoloration had not been reported up to now.
We chose two methods to achieve this goal: enzymatic discoloration with hydrogen peroxide, and
adsorption on a synthetic resin.
The chemical name of Direct Blue 273 is cuprate(3-),[3-[2-[5-(hydroxy-kO)-4-[2-[1-(hydroxy-
kO)-6-(phenylamino)-3-sulfo-2-naphthalenyl]diazenyl-kN1]-2-methylphenyl]diazenyl]-1,5-
naphthalenedisulfonato(5-)]-,sodium (1:3). The CAS number is 76359-37-0. The chemical
formula is C33 H20 CuN5 O11 S3 .3Na. The molecular weight is 891.26, and the structure is depicted
in Scheme 8.1.
The concentration of the dye in the commercial solution used in these experiments is 30%, and
the density of the solution is 1.034 g mL−1 . Some additives of this commercial solution are not
of public domain.

8.2 DB273 ENZYMATIC DISCOLORATION

8.2.1 The enzyme


Heme peroxidases have a ferriprotoporphyrin IX prosthetic group located at the active site (Dun-
ford, 1999). These enzymes catalyze the oxidation of a wide variety of substrates, using H2 O2 or
other peroxides by a mechanism that involves different intermediate enzyme forms. In the initial
step the native ferric enzyme is oxidized by hydrogen peroxide to form an unstable intermediate
called compound I (CoI), which has a structure of Fe IV = O-porphyrin π-cation radical, with
consequent reduction of peroxide to water. Then CoI oxidizes electron donor substrate to give com-
pound II (CoII), releasing a free radical. CoII is further reduced by a second substrate molecule,
regenerating the iron (III) state and producing another free radical. This difference between clas-
sical ping-pong and peroxidase ping-pong mechanisms is well documented, resulting from the
irreversible reactions of the peroxidase catalytic cycle (Everse et al., 1991).
Peroxidases from different sources have been successfully employed in removal of phenolic
pollutants, discoloration of synthetic dyes, in the paper pulp industry, in analysis and diagnostic
kits (Hamid and Khalil-ur-Rehman, 2009; Husain, 2010; Husain et al., 2009). Horseradish
peroxidase (HRP) isoenzyme C has been by far the most researched peroxidase. However, due to
134 P.Y.G. Clar et al.

Figure 8.1. Solutions of 0.16 µM DB273 in contact with 18 U mL−1 SBP, at different H2 O2 concentrations
(mM) presented at the top of the figure.

its high cost, much effort has been devoted to finding an alternative and peroxidases from several
sources have been investigated.
Soybean peroxidase (SBP) is an attractive alternative present at high concentration in the seed
coat, which is an inexpensive source, by-product of food industry (Gillikin and Graham, 1991;
Kamal and Behere, 2003). This is an anionic, haem-containing glycoprotein of 40662 Da with
an isoelectric point of 4.1. It can be extracted with a buffer solution and the coat could then be
pelleted for animal feed.
SBP shows an unusually high thermal and chemical stability even under acidic conditions. It was
successfully employed in removal of phenolic compounds and polycyclic aromatic hydrocarbons
from wastewater, and for bromination reactions (McEldoon et al., 1995).
The soybean seed coat peroxidase (SBP) can be extracted with a buffer solution and the coat
can then be pelleted for animal feed. This enzyme is ideal for environmental applications due
to the high catalytic efficiency even in its crude form, avoiding the cost of purifying the protein
(Steevensz et al., 2009).
The crude extract of SBP was obtained by contact of seed coats with a solution containing 50
sodium acetate/acetic acid and 5 mM CaCl2 (Magri et al., 2005). The homogenate was centrifuged
at 2000 g for 15 min and the supernatant was separated. The mass of protein in the clarified
homogenate was 50 ppm and the catalytic activity of the enzyme was 300 U mL−1 , determined
by comparison with the standard purchased from Sigma: Peroxidase from Glycine max (soybean)
P-1432, 50–150 units per mg of solid. One unit will form 1.0 mg of purpurogallin from pyrogallol
in 20 seconds at pH 6.0 at 20◦ C.
SBP was mostly employed for the oxidation of phenolic compounds, but there are very
few examples of applications on dye oxidation: the removal of the 90% of a soluble Cu(II)-
phthalocyanine (Marchis et al., 2011), and the 40% of the recalcitrant azo-dye for paper Direct
Yellow 11 (Knutson et al., 2005).

8.2.2 Color removal by oxidation


The crude extract of SBP was then proved together with hydrogen peroxide for DB273
discoloration.
The experiments for removal of DB273 were carried out in distilled water. The near-neutral
pH condition is preferable because it can easily apply to most wastewater.
Figures 8.1 and 8.2 exhibit different solutions of DB273 treated with H2 O2 in presence of SBP.
At H2 O2 concentration levels below 0.06 mM there is no evidence of dye degradation. On the
other hand, over 20 mM H2 O2 there is a loss of efficiency in color removal by the enzymatic way.
The kinetic experiments were carried out by a batch method (100 mL) at room temperature
(24◦ C) under constant stirring. The nominal enzyme concentration was varied between 0.3 and
9 U mL−1 in different experiments to determine the optimal conditions and to have an insight of
the multiple paths involved in the removal of this dye (Fig. 8.3).
The initial concentration of hydrogen peroxide was set in 0.20 mM for an initial dye concen-
tration of 87 ± 1 µM. This choice was based on our preliminary results and also on the optimal
hydrogen peroxide to substrate molar ratio described for other complex wastewater samples
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 135

Figure 8.2. Solutions of 0.16 µM DB273 in contact with 18 U mL−1 SBP, using 0.00, 0.20 and
0.06 mM H2 O2 .

Figure 8.3. Kinetics of enzymatic degradation of DB273 at 0.2 mM H2 O2 and different SBP activities: 0.0
(A), 0.3 (B), 0.6 (C), 1.2 (D), 2.4 (E), 3.0 (F) and 9.0 (G) U mL−1 .

(Nicell and Wright, 1997). It is well known that the addition of hydrogen peroxide in excess
causes inhibition by three known suicide pathways (Steevensz et al., 2009).
The discoloration kinetics was determined by following the decrease of dye concentration in
the solution by UV-visible spectrophotometry, measuring the absorbance at 588 as a function of
time during one hour. The solution of DB273 exhibited discoloration by the combined effect of
H2 O2 and SBP (Figs. 8.3 and 8.4).
In absence of the enzymatic extract, the color removal with H2 O2 was minimal (0 U mL−1 ,
curve A of Fig. 8.3). On the other hand, the enzymatic extract did not produce dye degradation
by itself, in absence of hydrogen peroxide.
Besides, when the SBP extract was previously boiled during two hours, the discoloration of
the solution of DB273 by action of H2 O2 was negligible due to the denaturation of the enzyme.
The nominal enzyme concentration was varied in different kinetic experiments (Fig. 8.3). For
the highest enzymatic activity tested (9.0 U mL−1 , curve G), the DB273 concentration decayed
to the 52% of the initial value after 1 minute of reaction, and to the 25% after 30 minutes.
136 P.Y.G. Clar et al.

Figure 8.4. UV-visible spectra of DB273 solution in contact with 0.6 U mL−1 SBP and 0.2 mM H2 O2 at
different times of reaction.

Table 8.2. Kinetic parameters for DB273 discoloration with 0.2 mM H2 O2 and with SBP at different activity
levels.
SBP activity
Parameters 0.3 U mL−1 0.6 U mL−1 1.2 U mL−1 2.4 U mL−1 3.0 U mL−1

[DB273]∞ 52.50 ± 0.29 41.30 ± 0.10 34.60 ± 0.11 22.20 ± 0.19 21.40 ± 0.12
(µM)
[DB273]1 22.8 ± 9.3 18.79 ± 0.72 34.10 ± 0.43 43.60 ± 0.60 41.70 ± 0.50
(µM)
k1 (min−1 ) 0.0589 ± 0.0077 0.3544 ± 0.0223 0.546 ± 0.013 1.157 ± 0.032 1.556 ± 0.039
[DB273]2 12.9 ± 9.1 16.30 ± 0.70 16.80 ± 0.35 20.40 ± 0.24 21.70 ± 0.21
(µM)
k2 (min−1 ) 0.0294 ± 0.0095 0.0683 ± 0.0023 0.0638 ± 0.0023 0.0474 ± 0.0016 0.0658 ± 0.0016

The concentration of DB273 in water was monitored as a function of time, and the experimental
data of curves B-F of Figure 8.3 were fit to the following empirical model:

[DB273]t = [DB273]€ + [DB273]1 × e−kt ×t + [DB273]2 × e−k2 ×t (8.1)

where [DB273]∞ is the amount of non-degraded dye, [DB273]1 and [DB273]2 the amounts
removed by two different processes, k1 and k2 the pseudo-first order kinetic constants for the
exponential decay of the dye concentration. The estimated parameters for the tested SBP activities
are presented in Table 8.2, Figures 8.5 and 8.6.
The experimental data presented in Figure 8.3 properly fitted this empirical model, except
those from an enzymatic activity of 9.0 U mL−1 (curve G).
The amount of dye that remained in solution when the equilibrium was reached ([DB273]∞ )
decayed exponentially with the nominal enzymatic activity (Fig. 8.5). The remaining color at high
enzymatic activity was attributed to other additives of the commercially available dye preparation
that are not substrate of this enzyme, or the reaction products.
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 137

Figure 8.5. Evolution of the kinetic parameters [DB273]∞ (•), [DB273]1 (◦) and [DB273]2 () for removal
of the anionic dye, as a function of the enzymatic activity.

Figure 8.6. Evolution of the kinetic parameters k1 (•) and k2 (◦) for removal of the anionic dye, as a function
of the enzymatic activity.
138 P.Y.G. Clar et al.

The parameters [DB273]1 and k1 from the second term of the empirical model also showed a
dependence with the SBP activity: k1 increased linearly (Fig. 8.6), and [DB273]1 rose to 42 µM
which represents a 48% of the total amount of the dye in solution (Fig. 8.5). This last parameter
showed a growing tendency, with an estimated maximum of 52 µM, indicating that almost the
60% of the dye could eventually be discolorized by this mechanism at higher amounts of enzyme
(Fig. 8.5). Besides, the analysis of the evolution of k1 showed a minimum of 0.096 U mL−1 of
SBP to observe color removal by this way (Fig. 8.6).
The parameter [DB273]2 from the third term of the empirical model also showed a dependence
with the SBP activity rising up to 22 µM, a 25% of the total amount of dye (Fig. 8.5). Instead,
the kinetic parameter k2 did not present any tendency with the enzymatic activity, and reached an
average value of 0.055 ± 0.016 min−1 (Fig. 8.6).
The analysis of these results indicated a color removal by enzymatic oxidation ruled by the
second term of the model.
The third term is indirectly related to the enzymatic activity. Here, the radical products of DB273
coming from the enzymatic oxidation would polymerize in solution destroying the chromophoric
system: the higher amount of radicals, the higher dye removal by this non-enzymatic path.
The equilibrium of this discoloration was reached after 60 minutes, with an efficiency of 73%
for an enzymatic activity of 3.0 U mL−1 .
Enzymatic oxidation of azo-dyes has been studied in depth, and we found similarity between the
efficiency of the system described and the results informed in literature. Some authors reported
color removal from azo-dyes solutions with efficiencies between 62 and 84% with horseradish
peroxidase (Venkata Mohan et al., 2005), and 60–80% with a soluble peroxidase from Trichosan-
thes dioica which was improved to almost 100% in presence of redox mediators such as riboflavin
(Jamal et al., 2011).

8.3 DB273 DISCOLORATION BY ADSORPTION

Up to now, there is not a unique method for treatment of all wastewater from dyeing industry
that can be considered complete and inexpensive (Gupta and Suhas, 2009). Usually the effluent
treatment includes more than one process to achieve the desired water quality, and the selection
depends on the nature of the wastewater.
Adsorption has been found to be superior compared to other techniques for wastewater treat-
ment in terms of its efficiency, versatility and simplicity of design (Haque et al., 2011; Khaled
et al., 2009; Ramakrishna and Viraraghavan, 1997; Wawrzkiewicz, 2011). It has been of growing
importance due to the chemical and biological stability of dyestuffs and the growing need for
high quality treatment.
In this work, we also present the discoloration of aqueous solutions of DB273 by adsorption on
a synthetic resin. A polyampholyte named Poly-(EGDE-MAA-2MI) synthesized in one-pot step
and milled in microparticles (Lázaro Martínez et al., 2008a; Leal Denis et al., 2008) was tested
as adsorbent for this anionic azo-dye. The performance of this method was then compared to the
enzymatic discoloration.

8.3.1 Synthesis and characterization of the polyampholyte


The material was synthesized by mixing one milliliter of methacrylic acid (MAA), 2.0 mL of
ethylene glycol diglycidyl ether (EGDE), 1.0 g de 2-methylimidazole (2MI) and 4.0 mL of ace-
tonitrile. After that, 30 mg of benzoyl peroxide was added to the reaction, thermostatted at 60◦ C
for 24 h. The product (Poly(EGDE-MAA-2MI)) was milled and washed three times with dis-
tilled water (Lázaro Martínez et al., 2008a; Leal Denis et al., 2008). Scheme 8.2 shows the most
probable chemical structure for Poly(EGDE-MAA-2MI).
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 139

Scheme 8.2. Probable chemical structure of the polyampholyte (Poly(EGDE-MAA-2MI)).

The mechanism of synthesis and the chemical structure of this material were studied by means of
a variety of spectroscopic methods, and carefully described in previous reports (Lázaro Martínez
et al., 2008a; 2008b; 2012; Leal Denis et al., 2008).
Here we present the characterization of the polyampholyte production batch employed for the
experiments of DB273 adsorption.
The elemental analysis of this batch of Poly(EGDE–MAA–2MI) was: C, 48.69; H, 7.70 and
N, 5.54.
The isoelectric point (pI) for this material determined by measurements of zeta potential (ζ)
was 5.5. The apparent diameter of the particles was 64 ± 4 µm.
These particles have titrable carboxylate and 2MI residues, besides the hydroxyl groups
(Scheme 8.2). The pI calculated from the titrable functional groups (4.61) was lower than the
value obtained from ζ measurements. To explain this difference we must focus on the nature of
the ionic and ionizable sites. These particles have carboxylate groups (1.48 mmol g−1 ), hydroxyl
groups, and 2MI residues. These heterocycles can be: monosubstituted (2MI), which can also be
protonated (2MIH+ ); or disubstituted (2MI+ ) bearing a permanent positive electric charge. Only
R-CO2 H and 2MIH+ could be determined by titration of the acid form of this polyampholyte,
which brought to a theoretical pI lower than the experimental pI from ζ measurements.
The amount of 2MI+ (0.36 mmol g−1 ) was determined indirectly (Lázaro Martínez et al., 2012).
The total number of 2MI+ residues in the polymer deduced from the elemental analysis (1.98
mmol g−1 ) was compared with the number of titrable 2MI (0.83 mmol g−1 ) and of 2MI+ ). From
this analysis, we observed that 0.79 mmol g−1 of residues were not detected (“non-accessible
2MI”).

8.3.2 Kinetics of sorption


Adsorption experiments were conducted in distilled water in order to imitate environmental con-
ditions and to avoid changes that would make the wastewater treatment more expensive. At
pH 6.0, the zwitterion bears 0.83 mmol of 2MIH + g-1 and 0.36 mmol of 2MI + g-1, totalizing
1.19 mmol of positive binding sites per gram, even if the average charge of the polymeric unit is
−0.23 (Lázaro Martínez et al., 2012).
Kinetic studies of DB273 removal by adsorption on the polyampholyte were carried out with
0.1000 of the polyampholyte in contact with 200 mL of 70 µM DB273 solution, at 24◦ C. Dye
concentration in the supernatant solution was monitored spectrophotometrically at 588 nm as a
function of time (Fig. 8.7).
Experimental data of DB273 concentration as a function of time fitted well to the pseudo-first
order rate expression of Lagergren, with a regression coefficient (R) of 0.9993 and a kinetic
140 P.Y.G. Clar et al.

80

60
[DB273] (uM)

40

20

0
0 10 20 30
Time (min)

Figure 8.7. Evolution of [DB273] by contact with the polyampholyte.

constant (k) of 0.1324 ± 0.0016 min−1 . After 30 minutes of contact the dye remaining in solution
was less than 3%.
This model quantitatively describes a sorption preceded by diffusion through a boundary layer.
The high correlation indicated that mass transport was the rate-limiting step, and intraparticle
diffusion could be neglected, in agreement with previous reports for this material (Copello et al.,
2012).

8.3.3 Isotherm data analysis


Isotherms were performed at 4, 18 and 24◦ C (Fig. 8.8), with 0.0500 g of polyampholyte and
100 mL of DB273 solution in a 20–220 µM concentration range. Samples were centrifuged and
filtered after 48 hours of contact time to reach the equilibrium. Free DB273 concentration in each
supernatant solution was determined spectrophotometrically at 588 nm.
The data of DB273 sorption on the polyampholyte at different temperatures were examined by
using four types of the most common isotherms: Langmuir, Freundlich, Dubinin–Radushkevich
(D–R), and Temkin isotherm models. The estimated parameters for the experiment performed at
4, 18 and 24◦ C are presented in Tables 8.3, 8.4 and 8.5, respectively.
The best model varied according with temperature. At 4◦ C and at 24◦ C the best fit was obtained
by means of the Temkin isotherm (Eq. 8.2), closely followed by Langmuir isotherm of one type
of site (Eq. 8.3), according with the evidence ratio (ER) based on the corrected Akaike criterion
(AIC c ; Tables 8.3 and 8.5) (Copello et al., 2012).

RT
qe = ln(Ce kT ) (8.2)
b
qm Ce
qe = (8.3)
Ce + K L
where Ce is the concentration of dye at equilibrium, qe is the adsorption capacity in equilibrium
with the corresponding Ce , T is the absolute temperature in Kelvin, R is the universal gas constant
(0.008314 kJ mol−1 K−1 ), kT is the equilibrium binding constant, b is corresponding to the heat
of sorption, qm is the maximum adsorption capacity describing a complete monolayer adsorption,
and KL is the equilibrium constant of dissociation.
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 141

Figure 8.8. Adsorption isotherms of DB273 onto the polyampholyte at 4◦ C (), 18◦ C (♦) and 24◦ C ().

Table 8.3. Isotherm parameters for DB273 adsorption at 4◦ C on the polyampholyte. Regression coefficients,
values of AIC c criterion and evidence ratio (ER) for different models.

Model Parameters for T : 277 K R2 AIC c ER

Langmuir qm KL G 0.9861 56.74 1


335 ± 24 mg g−1 36 ± 5 mg L−1 −23.3 ± 1.5
kJ mol−1
Dubinin- qm KDR E 0.9763 63.25 25.9
Radushkevich 2.94 × 103 5.30 × 10−3 9.71 ± 0.37
± 0.56 × 103 ±0.40 × 10−3 kJ mol−1
mg g−1 mol2 kJ−2
Freundlich KF n 0.9681 66.72 147
6.4 × 104 1.67 ± 0.13
± 2.9 × 104
Temkin KT b 0.9859 56.94 1.1
0.371 ± 0.027 0.0351
L mg−1 ± 0.0013
kJ mol−1

qe (mg g−1 ) vs. Ce (mg L−1 ) except for D-R model in which Ce is in g g−1 .

The Temkin model suggests an equal distribution of binding energies over the number of sites
on the surface. It assumes that the heat of adsorption of all the molecules in layer decreases
linearly with coverage due to adsorbent-adsorbate interactions.
In the same way, the Langmuir model suggests that the sorption onto the surface is homogeneous
in nature.
At 18◦ C, all the tested models had statistical support and the best fit was obtained with
Freundlich equation (Table 8.4).
142 P.Y.G. Clar et al.

Table 8.4. Isotherm parameters for DB273 adsorption at 18◦ C on the polyampholyte. Regression
coefficients, values of AIC c criterion and evidence ratio (ER) for different models.

Model Parameters for T : 291 K R2 AIC c ER

Langmuir qm KL G 0.9733 63.05 1.5


482 ± 68 31.0 ± 7.9 −24.8 ± 1.3
mg g−1 mg L−1 kJ mol−1
Dubinin- qm KDR E 0.9737 63.78 2.1
Radushkevich 6.2 × 103 ± 5.2 × 10−3 9.81 ± 0.38
1.5 × 103 ± 0.4 × 10−3 kJ mol−1
mg g−1 mol2 kJ−2
Freundlich KF n 0.9712 62.29 1
23.1 ± 3.8 1.48 ± 0.11
Temkin KT b 0.9583 67.74 15.3
0.531 ± 0.064 0.0289 ± 0.0020
L mg−1 kJ mol−1

qe (mg g−1 ) vs. Ce (mg L−1 ) except for D-R model in which Ce is in g g−1 .

Table 8.5. Isotherm parameters for DB273 adsorption at 24 ◦ C on the polyampholyte. Regression
coefficients, values of AIC c criterion and evidence ratio (ER) for different models.

Model Parameters for T : 297 K R2 AIC c ER

Langmuir qm KL G 0.9714 53.24 1.93


403 ± 44 17.9 ± 4.1 −26.7 ± 2.7
mg g−1 mg L−1 kJ mol−1
Dubinin- qm KDR E 0.9531 57.71 18.13
Radushkevich 4.3 × 103 4.30 × 10−3 10.78 ± 0.63
±1.4 × 103 ±0.50 × 10−3 kJ mol−1
mg g−1 mol2 kJ−2
Freundlich KF n 0.9444 59.24 38.86
33.8 ± 7.3 1.69 ± 0.20
Temkin KT b 0.9753 51.92 1
0.613 ± 0.069 0.0289 ± 0.0017
L mg−1 kJ mol−1

qe (mg g−1 ) vs. Ce (mg L−1 ) except for D-R model in which Ce is in g g−1 .

The average adsorption capacity predicted by the Langmuir isotherm model was in the level of
0.46 ± 0.08 mmol g−1 , occupying a 39% of the positive binding sites on the polymeric surface.
The efficiency for color removal by adsorption on the polyampholyte was 88%, when the
initial concentration of DB273 was 87 µM. On the other hand, the equilibrium of enzymatic
discoloration was reached after 60 minutes for the same initial concentration of DB273, with an
efficiency of 73% for an enzymatic activity of 3.0 U −1 . Thus, we can say that these two selected
methods exhibited a similar performance.
This predicted loading capacity from Langmuir isotherm (0.46 ± 0.08 mmol g−1 ) was higher
compared with the adsorption of other anionic and cationic dyes on inorganic sorbents reported in
literature (Khaled et al., 2009; Lin et al., 2011; Tehrani-Bagha et al., 2011), and lower compared
with a commercially available ion-exchange resin (Wawrzkiewicz, 2011). On the other hand, the
equilibrium constant of dissociation (KL , expressed in M units) was lower than the values reported
in all the mentioned examples.
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 143

Table 8.6. Saline effect on DB273 adsorption on the polyampholyte surface.

System [DB273]* (µM)

DB273 in water 124.8


polyampholyte + DB273 in water 20.0
polyampholyte + DB273 in 1 M KNO3 36.5
polyampholyte + DB273 in 0.2 M KNO3 38.9
polyampholyte + DB273 in 0.05 M KNO3 44.4

* At equilibrium

Values of G of adsorption obtained at the three temperatures were close to −20 kJ mol−1 ,
indicating that physisorption was the predominant binding mechanism (Tables 8.3, 8.4 and 8.5)
(Tehrani-Bagha et al., 2011). The increase of T enhanced the spontaneity of the adsorption,
producing a shift of G to more negative values.
Enthalpy and entropy of adsorption are given from van’t Hoff plots. The H value was
21 kJ mol−1 and the S value was 0.16 kJ mol−1 . The positive value of H indicated that the
adsorption was endothermic in nature. Typically, the enthalpy of a physisorption process is lower
than 40 kJ mol−1 , attributed to a number of physical interactions involving electrostatic attrac-
tions, non-polar interactions, water bridging, hydrogen-type bonding, and ion-exchange reactions
between dye molecules and the adsorbent structures (Cestari et al., 2007). Here, sulfonate residues
from the dye were expected to interact with the protonated 2MI residues (2MIH+ ) and with the
fixed positive charges from the disubstituted 2MI units in the network (2MI+ ), but at the same
time the presence of carboxylate groups from the polyampholyte should weaken this binding by
electrostatic repulsion.
The positive value of S indicated that there was an increase in the randomness in the system
solid/solution interface with release of water molecules from the surface into the solution (Sevim,
2011).
The Dubinin–Radushkevich isotherm (Eq. 8.4), assumes a constant sorption potential:
   2
1
ln qe = ln qm − KDR RT ln 1 + (8.4)
Ce
The magnitude of the mean free energy of sorption (Tables 8.3, 8.4 and 8.5 and Eq. 8.5) is
related to the type of sorption reaction: a range from about 8 to 16 kJ mol−1 indicates ion-exchange
processes. The value close to 10 kJ mol−1 indicated that the dye adsorption was controlled by both
weak physical forces and, mainly, water adsorbed/dye in solution ion-exchange interactions:

1
E= √ (8.5)
2 KDR
The effect of salt addition on dye adsorption was studied from a system containing 120 mL of
125 µM of DB273 and KNO3 at different concentration levels, and 0.0600 g of polyampholyte
(Table 8.6).
The amount of Direct Blue 273 retained by this sorbent decreased in a range between 16 and
23% upon the addition of KNO3 , when compared with the adsorption of the dye dissolved in
pure water. Competition of NO3 - with the anionic dye for the positive binding sites on the surface
was less significant at the highest concentration level tested (1 M), attributed to a lower ionic
activity.
The effect of ionic species on the dye desorption from the polymeric surface was explored using
0.0740 g of dye-loaded polyampholyte in 120 mL of solution. Sodium dodecyl sulfate (SDS), and
KNO3 were selected as ionic species. Desorption of DB273 from the loaded polyampholyte was
144 P.Y.G. Clar et al.

Table 8.7. Saline effect on DB273 desorption from the


loaded polyampholyte.

System [DB273]* (µM)

DB273-polyampholyte + water 0
DB273-polyampholyte + 1 M KNO3 0
DB273-polyampholyte + 1 mM SDS 0.4
DB273-polyampholyte + 80 mM SDS 13.4

* At equilibrium

Figure 8.9. Images of the polyampholyte milled in particles (a), the polyampholyte loaded with DB273 (b),
and the DB273-polyampholyte treated with 0.10 M NaClO (c).

negligible with 1 KNO3 and with 1 mM SDS. A solution of 80 mM SDS produced the release of
a 10% of the adsorbed amount of DB273 (Table 8.7).
Finally, 0.0740 g of dye-loaded polyampholyte were treated with 120 mL of solutions of sodium
hypochlorite as oxidizing agent during 4 hours at two concentration levels: 0.10 and 0.35 M.
Discoloration by oxidation was effective in both cases (Fig. 8.9), but treatment with 0.35 M
sodium hypochlorite produced partial degradation of the adsorbent.

8.3.4 FTIR analysis


Figure 8.10 shows the FTIR spectra for Poly(EGDE-MAA-2MI), for the dye DB273 and for the
DB273-loaded polyampholyte.
Complete FTIR analysis of the polymer has been previously reported (Denaday et al., 2011;
Lázaro Martínez et al., 2008a; 2008b; 2012; Leal Denis et al., 2008). Bands at 1721, 1650/1450,
1570 and 1100 cm−1 were attributed to the stretching vibration of C=O, C=C/N (heterocyclic
rings), R-CO− 2 and C-O respectively. Bending vibrations of the 2MI residue and its torsion
stretching were present at around 857, and at 757 and 672 cm−1 , respectively.
Dye’s FTIR spectrum analysis put in evidence the presence of N=N bond from stretching
vibrations at 1591 and 1615 cm−1 , and the sulfonate groups from SO bond vibration bands at
1183, 1032 and 611 cm−1 . Bands at 1615 and 1591 cm−1 were attributed to N-H bending, and
the band at 1300 cm−1 to the stretching of C-N (Coates, 2000).
The spectrum of the dye-loaded polyampholyte predominantly shows the adsorbent’s bands,
even if the dye adsorption on the polymeric surface is evident (Fig. 8.9b).
Only two significant spectral changes can be mentioned as a consequence of the adsorption.
The 2MI torsion stretching bands at 757 and 672 cm−1 were replaced by a new band between
670 and 615 cm−1 . This change would be indicative of electrostatic interaction of 2MIH+ ) and
2MI+ residues with the R-SO− 3 from the dye, or either some interaction between 2MI residues
and the Cu(II) center in the DB273 molecule. Besides, a R-CO− 2 band broadening at 1570 cm
−1

could be attributed to a simple overlapping of bands from the two interacting species, or instead
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 145

Figure 8.10. FTIR spectra of Poly(EGDE-MAA-2MI) (A), DB273 (B) and DB273-loaded poly(EGDE-
MAA-2MI) (C).

to electrostatic repulsion with R-SO−


3 , or some interaction between carboxylate and the Cu(II)
center of DB273.

8.4 CONCLUSIONS

This was the first attempt of DB273 discoloration reported in literature up to now.
The crude extract of SBP turned out to be a powerful biotechnological tool for the ecofriendly
dye degradation, using hydrogen peroxide as oxidant. Up to now, this enzyme had only been
tested on a few dyes, such as the Cu(II)-phthalocyanine Remazol Turquoise Blue G133 (Marchis
et al., 2011), the azo dye Direct Yellow 11 and the methine dye Basazol 46L (Knutson et al.,
2005). The discoloration efficiency of SBP on DB273 was strongly dependent on the amount of
crude extract employed. An activity of 3.0 U mL was necessary to remove the 73% of the dye in
60 minutes, using 0.2 mM H2 O2 for a 87 µM initial concentration of dye. A similar efficiency
could be reached in 30 minutes with an activity of 9.0 U mL−1 . Results were comparable to those
obtained with peroxidases from other sources on different azo dyes, but the cost of the process
with SBP would be significantly lower.
The polyampholyte was also an effective adsorbent for the anionic dye at near-neutral pH.
The rate of sorption was controlled by the diffusion through a boundary layer. The Temkin and
Langmuir models fitted well the results from the equilibrium of physisorption, predicting that
adsorption energies of the active sites were equivalent on the polyampholyte surface. The load-
ing capacity was higher compared with inorganic sorbents, and somewhat lower compared with
commercially available ion-exchange resins. On the other hand, the equilibrium constant of disso-
ciation (KL , expressed in M units) was found to be lower, and the dissolved salts from the sample
matrix would not affect significantly the efficiency of uptake.
Discoloration of the dye-loaded polyampholyte was successfully achieved with 0.10M NaClO
as oxidizing agent, preserving the adsorbent’s chemical structure. The preconcentration of the dye
on the adsorbent surface allows to reduce the amount of oxidant necessary to achieve degradation.
146 P.Y.G. Clar et al.

ACKNOWLEDGEMENTS

The authors thank the financial support from Universidad de Buenos Aires (UBACyT 10-12/237),
CONICET (CONICET 2010-2012/PIP 100076), ANPCyT (BID 1728/PICT 01778 and PICT-
2010-1957). Pamela Yanina González Clar thanks the Universidad de Buenos Aires for her
research fellowship. The authors also thank Eng. Hugo Velez (INTI-Argentina) for the kind
supply of a commercial solution of the azo dye Direct Blue 273.

REFERENCES

Ahmed Basha C., Selvakumar K.V., Prabhu H.J., Sivashanmugam P. & Lee C.W.: Degradation studies for
textile reactive dye by combined electrochemical, microbial and photocatalytic methods. Sep. Purif.
Technol. 79:3 (2011), pp. 303–309.
Baig, S. & Liechti P.A.: A ozone treatment for biorefractory COD removal. Water Sci. Technol. 43:2 (2001),
pp. 197–204.
Cestari, A.R., Vieira, E.F., Vieira, G.S. & Almeida, L.E.: Aggregation and adsorption of reactive dyes in
the presence of an anionic surfactant on mesoporous aminopropyl silica. J. Colloid Interface Sci. 309:2
(2007), pp. 402–411.
Chung, K.T. & Cerniglia, C.E.: Mutagenicity of azo dyes: structure -activity relationships. Mutat. Res. 277:3
(1992), pp. 201–220.
Coates, J.: Interpretation of infrared spectra, a practical approach.In: R.A. Meyers (ed): Encyclopedia of
analytical chemistry. John Wiley & Sons Ltd, Chichester, England, 2000, pp. 10,815–10,837.
Cooper, P. (ed): Colour in dyehouse effluent. Society of Dyers and Colorists, Bradford, West Yorkshire BD1
2JB, 1995.
Copello, G.J., Diaz, L.E. & Campo Dall’ Orto, V.: Adsorption of Cd(II) and Pb(II) onto a one step-synthesized
polyampholyte: Kinetics and equilibrium studies. J. Hazard. Mater. 217/218 (2012), pp. 374–381.
Denaday, L.R., Miranda, M.V., Torres Sánchez, R.M., Lázaro Martínez, J.M., Lombardo Lupano, L.V. &
Campo Dall’ Orto, V.: Development and characterization of a polyampholyte-based reactor immobilizing
soybean seed coat peroxidase for analytical applications in a flow system. Biochem. Eng. J. 58/59 (2011),
pp. 57–68.
Dos Santos, A.B., Cervantes, F.J. & Van Lier, J.B.: Review paper on current technologies for decolourisation of
textile wastewaters: Perspectives for anaerobic biotechnology. Bioresour. Technol. 98:12 (2007), pp. 2369–
2385.
Dunford, H.B.: Heme peroxidases. John Wiley and Sons, New York, NY, USA, 1999.
Elias B., Guihard L., Nicolas S., Fourcade F. & Amrane A.: Effect of electro-Fenton application on azo dyes
biodegradability. Environ. Prog. Sust. Energy 30:2 (2011), pp. 160–167.
Everse, J., Everse, K. E., & Grisham, M. B. (eds): Peroxidases in chemistry and biology, Vol. 2. CRC Press,
Boca Raton, FL, USA, 1991.
Ferreira-Leitao, V.S., de Carvalho, M.E.A. & Bon, E.P.S.: Lignin peroxidase efficiency for methylene blue
discoloration: Comparison to reported methods. Dyes and Pigments 74:1 (2007), pp. 230–236.
Gillikin, J.W. & Graham, J.S.: Purification and developmental analysis of the major anionic peroxidase from
the seed coat of Glycine max. Plant Physiol. 96:1 (1991), pp. 214–220.
Gupta, V.K. & Suhas: Application of low-cost adsorbents for dye removal – A review. J. Environ. Manage.
90:8 (2009), pp. 2313–2342.
Hamid, M. & Khalil-ur-Rehman: Potential applications of peroxidases. Food Chem. 115:4 (2009), pp.
1177–1186.
Haque, E., Jun, J.W. & Jhung, S.H.: Adsorptive removal of methyl orange and methylene blue from aqueous
solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard. Mater. 185:1
(2011), pp. 507–511.
Husain Q.: Peroxidase mediated decolorization and remediation of wastewater containing industrial dyes: A
review. Rev. Environ. Sci. Biotechnol. 9:2 (2010), pp. 117–140.
Husain, Q., Husain, M. & Kulshrestha, Y.: Remediation and treatment of organopollutants mediated by
peroxidases: A review. Crit. Rev. Biotechnol. 29:2 (2009), pp. 94–119.
Jamal, F., Qidwai, T., Pandey, P.K., Singh, R. & Singh, S.: Azo and anthraquinone dye decolorization in
relation to its molecular structure using soluble Trichosanthes dioica peroxidase supplemented with redox
mediator. Catal. Commun. 12:13 (2011), pp. 1218–1223.
Eco-friendly approach for Direct Blue 273 removal from an aqueous medium 147

Kamal, J.K. & Behere, D.V.: Activity, stability and conformational flexibility of seed coat soybean peroxidase.
J. Inorg. Biochem. 94:3 (2003), pp. 236–242.
Khaled, A., El Nemr, A., El-Sikaily, A. & Abdelwahab, O.: Removal of Direct N Blue-106 from artificial
textile dye effluent using activated carbon from orange peel: Adsorption isotherm and kinetic studies.
J. Hazard. Mater. 165:1–3 (2009), pp. 100–110.
Knutson, K., Kirzan, S. & Ragauskas, A.: Enzymatic biobleaching of two recalcitrant paper dyes with
horseradish and soybean peroxidase. Biotechnol. Lett. 27:11 (2005), pp. 753–758.
Lázaro Martínez, J.M., Leal Denis, M.F., Campo Dall’Orto, V. & Buldain, G.Y.: Synthesis, FTIR, solid-state
NMR and SEM studies of novel polyampholytes or polyelectrolytes obtained from EGDE, MAA and
imidazoles. Eur. Polym. J. 44:2 (2008a), pp. 392–407.
Lázaro Martínez, J.M., Chattah, A.K., Monti, G.A., Leal Denis, M.F., Buldain, G.Y. & Campo Dall’ Orto,
V.: New copper(II) complexes of polyampholyte and polyelectrolyte polymers: Solid-state NMR, FTIR,
XRPD and thermal analyses. Polymer 49:25 (2008b), pp. 5482–5489.
Lázaro Martínez, J.M., Chattah, A.K., Torres Sánchez, R.M., Buldain, G.Y. & Campo Dall’Orto, V.: Synthesis
and characterization of novel polyampholyte and polyelectrolyte polymers containing imidazole, triazole
or pyrazole. Polymer 53:6 (2012), pp. 1288–1297.
Leal Denis, M.F., Carballo, R.R., Spiaggi, A.J., Dabas, P.C., Campo Dall’ Orto, V., Lázaro Martínez, J.M. &
Buldain, G.Y.: Synthesis and sorption properties of a polyampholyte. React. Funct. Polym. 68:1 (2008),
pp. 169–181.
Lin, Y.F., Chen, H.W., Chien, P.S., Chiou, C.S. & Liu, C.C.: Application of bifunctional magnetic adsorbent
to adsorb metal cations and anionic dyes in aqueous solution. J. Hazard. Mater. 185:2–3 (2011), pp. 1124–
1130.
McEldoon, J.P., Pokora, A.R. & Dordick, J.S.: Lignin peroxidase-type activity of soybean peroxidase. Enzyme
Microb. Technol. 17:4 (1995), pp. 359–365.
Magri, M.L., Miranda, M.V. & Cascone, O.: Immobilization of soybean seed coat peroxidase on polyaniline:
Synthesis optimization and catalytic properties. Biocatal. Biotransform. 23:5 (2005) pp. 339–346.
Majewska-Nowak, K., Winnicki, T. & Wiśniewski, J.: Effect of flow conditions on ultrafiltration efficiency
of dye solutions and textile effluents. Desalination 71:2 (1989), pp. 127–135.
Marchis, T., Avetta, P., Bianco-Prevot, A., Fabbri, D., Viscardi, G. & Laurenti, E.: Oxidative degra-
dation of Remazol Turquoise Blue G 133 by soybean peroxidase. J. Inorg. Biochem. 105:2 (2011),
pp. 321–327.
Mohajerani M., Mehrvar M. & Ein-Mozaffari F.: An overview of the integration of advanced oxidation
technologies and other processes for water and wastewater treatment. Int. J. Eng. 3:2 (2009), pp. 120–146.
Nicell, J.A. & Wright, H.: A model of peroxidase activity with inhibition by hydrogen peroxide. Enzyme
Microb. Technol. 21:4 (1997), pp. 302–310.
Oller, I., Malato, S. & Sánchez-Pérez, J.A.: Combination of advanced oxidation processes and biological
treatments for wastewater decontamination - A review. Sci. Total Environ. 409:20 (2011), pp. 4141–4166.
Park, C., Lee, M., Lee, B., Kim, S.W., Chase, H.A., Lee, J. & Kim, S.: Biodegradation and biosorption for
decolorization of synthetic dyes by Funalia trogii. Biochem. Eng. J. 36:1 (2007), pp. 59–65.
Ramakrishna, K. & Viraraghavan, T.: Dye removal using low cost adsorbents. Water Sci. Technol. 36:2–3
(1997), pp. 189–196.
Ruppert, G., Bauer, R. & Heisler, G.: UV-O3 , UV-H2 O2 , UV-TiO2 and the photo-Fenton reactions comparison
of advanced oxidation processes for wastewater treatment. Chemosphere 28:8 (1994), pp. 1447–1454.
Sevim, A.M., Hojiyev, R., Gül, A. & Çelik, M.S.: An investigation of the kinetics and thermodynamics of
the adsorption of a cationic cobalt porphyrazine onto sepiolite. Dyes Pigments 88:1 (2011), pp. 25–38.
Solis, M., Solis, A., Perez, H. I., Manjarrez, N. & Flores, M.: Microbial discoloration of azo dyes: A review.
Process Biochem. 47:12 (2012), pp. 1723–1748.
Srinivasan, A. & Viraraghavan, T.: Decolorization of dye wastewaters by biosorbents: A review. J. Environ.
Manage. 91:10 (2010), pp. 1915–1929.
Steevensz, A., Mousa Al-Ansari, M., Taylor, K.E., Bewtra, J.K. & Biswas, N.: Comparison of soybean
peroxidase with laccase in the removal of phenol from synthetic and refinery wastewater samples. J.
Chem. Technol. Biotechnol. 84:5 (2009), pp. 761–769.
Tehrani-Bagha, A.R., Nikkar, H., Mahmoodi, N.M., Markazi, M. & Menger, F.M.: The sorption of cationic
dyes onto kaolin: Kinetic, isotherm and thermodynamic studies. Desalination 266:1–3 (2011), pp. 274–
280.
Torres, E., Bustos-Jaimes, I. & Le Borgne, S.: Potential use of oxidative enzymes for the detoxification of
organic pollutants. Appl. Catal. B 46:1 (2003), pp. 1–15.
148 P.Y.G. Clar et al.

Venkata Mohan, S., Krishna Prasad, K., Chandrasekhara Rao, N. & Sarma, P.N.: Acid azo dye degradation
by free and immobilized horseradish peroxidase (HRP) catalyzed process. Chemosphere 58:8 (2005), pp.
1097–1105.
Wawrzkiewicz, M.: Comparison of gel anion exchangers of various basicity in direct dye removal from
aqueous solutions and wastewaters. Chem. Eng. J. 173:3 (2011), pp. 773–781.
Xiaobin, C., Rong, J., Pingsheng, L., Shiqian, T., Qin, Z., Wenzhong, T. & Xudong, L.: Purification of a
new manganese peroxidase of the white-rot fungus Schizophyllum sp. F17, and decolorization of azo dyes
by the enzyme. Enzyme Microb. Technol. 41:3 (2007), pp. 258–264.
CHAPTER 9

Decontamination of commercial chlorpyrifos in water


using the UV/H2 O2 process

Joana Femia, Melisa Mariani, Alberto Cassano, Cristina Zalazar & Inés Tiscornia

9.1 INTRODUCTION

In the last few years, both the extensive use of pesticides to increase agricultural production and
the intensive development of new chemicals have dramatically increased the variety and amount
of agrochemicals present in the environment. In Argentina, for example, the use of glyphosate as
a herbicide has increased from 1 million liters in 1991 to 180 million liters in 2007 (Binimelis
et al., 2009). Chlorpyrifos is the most widely used insecticide. In 2010, more than 8 million liters
were used, which represent an increase of 100% compared with the number registered in 2006
(SENASA, Argentina).
Pesticide wastes from containers and rinsing equipments constitute a serious problem in many
countries. The inappropriate disposal of such wastes causes the contamination of soil, groundwater
and surface waters. Consequently, it is necessary to develop a treatment to remove these harmful
substances from the environment.
Chlorpyrifos (CP, o,o-diethyl-o-3,5,6-trichloro-2-pyridylphosphorothioate; Fig. 9.1) is an
organophosphate pesticide which is commercially supplied dissolved in organic solvents with
emulsifying additives owing to its low solubility in water (1.39 mg L−1 at 25◦ C, Giesy et al.,
1999). Emulsifiable concentrate formulations represent the largest volume of all formulations
deployed worldwide (Knowles, 2008; Kundu et al., 2012). They are used for insect control in
several crops, such as cotton, corn and alfalfa. Also CP emulsifiable formulations are used as
anti-parasitics on cattle. CP is moderately persistent in soils with a medium life of less than 1 day
to 240 days, depending on the soil type (Pino and Peñuela, 2011).
As most organophosphates, when this pesticide is used incorrectly or excessively, it may cause
significant environmental damage and, at the same time, it could generate DNA damage, cell
malformations, and neurological damage in humans and animals. CP was also considered as
the most toxic insecticide for two marine invertebrates (Artemia sp. and Brachionous plicatilis)
(Guzzella et al., 1997). In 2009, Argentina’s Ministry of Health banned CP for domestic use,
since it was considered a threat to public health (Res. N◦ 456).
Different methods of CP degradation have been presented in recent papers. Zhang et al. (2011)
reported ultrasonic irradiation as an attractive technique for the degradation of chlorpyrifos and
diazinon in aqueous medium without the addition of other chemicals. Biodegradation is consid-
ered a viable and environmentally friendly approach for the decontamination of CP mainly in soil
(Briceño et al., 2012). A number of studies have shown that dissolved metal ions can play an
important role in the catalysis of organophosphorus pesticides, for example enhancing the rate of

Figure 9.1. Molecular structure of CP.

149
150 J. Femia et al.

hydrolysis reactions to achieve the pollutant degradation. Sarkouhi et al. (2012) investigated the
catalytic effect of Ag+ ions on the hydrolysis of chlorpyrifos and phoxim.
Among the technologies to degrade pesticides, advanced oxidation processes (AOPs) are an
interesting option for this type of pollutants (Ikehata and El-Din, 2006). They are based on
the generation of highly oxidizing transient species (especially the hydroxyl radical HO• ) and
include technologies such as O3 /H2 O2 , photo-Fenton, O3 /UV, H2 O2 /UV, and other photocatalytic
processes. AOP are important technologies at low flow rates with medium or low organic loads
(approximately 50 m3 h−1 and ≤1000 mg L−1 TOC) (García Molina, 2006). Therefore, these pro-
cesses are appropriate to treat the wastewater from the rinsing of empty containers to be recycled.
Several articles deal with the removal of organophosphorus pesticides in water using AOPs.
Shemer and Linden (2006) studied kinetics and degradation products resulting from the appli-
cation of UV and UV/H2 O2 to diazinon; Badawy et al. (2006) investigated the degradation of
organophosphorus containing substrates such as fenitrothion, diazinon and profenofos using UV,
UV/H2 O2 , Fenton and the photo-Fenton treatment.
Measuring the disappearance of a contaminant in AOPs is not enough because the treatment
could produce a variety of organic intermediates, which can be more toxic than the parent com-
pound. For this reason, toxicity measurements are very important in AOPs. Ecotoxicity tests
usually provide important information of a possible hazard potential of a substance on repre-
sentative organisms, especially aquatic ones such as the luminescent marine bacterium (Vibrio
fischeri), water fleas (Daphnia) or algae (Zooglea) (Aydin et al., 2011; Boluda et al., 2002). The
Microtox test with Vibrio fischeri has been widely used for assessing the toxicity of environmental
samples (Kralj et al., 2007). This bioluminescence inhibition assay is often chosen as the first
test in a battery of tests because of their speed and low cost.
Few studies in the literature deal specifically with chlorpyrifos treatment using AOPs. Murillo
et al. (2010) studied CP (analytical reagent grade) degradation by using photo-Fenton, TiO2 ,
TiO2 /H2 O2 , O3 and O3 /H2 O2 processes, comparing the effectiveness of each treatment. Chioma
Affam et al. (2012) evaluated the effect of various operating conditions of the solar photo-Fenton
process on a solution of chlorpyrifos, cypermethrin and chlorothalonil. Gomathi Devi et al. (2009)
studied the photocatalytic activity of polycrystalline TiO2 doped with different transition metals,
under UV/solar light, using CP as probe molecule. Kundu et al. (2012) established that both CP
in a commercial emulsifiable concentrate and its toxic hydrolysis product 3,5,6-trichloropyridin-
2-ol could be easily eliminated via a tandem process employing a TAML activator (tetra-amido
macrocyclic ligand, a small synthetic molecule mimics of peroxidase and short-circuited P450
enzymes) with H2 O2 . This degradation procedure produced a 72.5-fold reduction in the toxicity
(Microtox® assays) of the treated reaction mixture. Wu and Linden (2010) discussed the degrada-
tion of parathion and chlorpyrifos (analytical grade reagents) using UV and UV/H2 O2 , studying
the role of HO• and CO•− 3 radicals in aqueous solution. They reported that the combination of
H2 O2 and UV radiation enhances the degradation rate of CP with respect to only UV radiation in
aqueous solution.
In this work, a commercial formulation of chlorpyrifos in water was treated using the UV/H2 O2
process. To the best of our knowledge, the degradation of commercial formulations of CP in water
by this method has not been so widely studied. The optimal amount of oxidizing agent to be incor-
porated and the behavior of the degradation rate of chlorpyrifos were studied. Total organic carbon
was measured in order to evaluate the mineralization rate of the treated mixture. In addition, the
toxicity during the photodegradation was evaluated by employing the Microtox acute toxicity test.

9.2 MATERIALS AND METHODS

9.2.1 Chemicals
The following reagents were used: (i) chlorpyrifos Pestanal (Riedel-de-Haën) as chromatographic
standard; (ii) chlorpyrifos emulsifiable concentrate, 48% (active ingredient) provided by Red
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 151

Figure 9.2. Schematic representation of the laboratory scale experimental device. (a) Annular photoreactor
design. (b) Experimental device.

Surcos (Santa Fe, Argentina); (iii) hydrogen peroxide (H2 O2 , Ciccarelli p.a., 30%); (iv) catalase
from bovine liver, >2000 units/mg (Fluka, 1 unit decomposes 1 mmol H2 O2 per minute at pH 7.0
and 25◦ C). Pure distilled water was used in all experiments.

9.2.2 Experimental setups and procedures


The photodegradation of the chlorpyrifos emulsion was carried out in an annular reactor (reactor
volume, VR = 1200 mL), designed with the inner tube of quartz to allow the passage of UV radi-
ation from a BTE lamp Model 6169 KV (λ = 253.7 nm, 20 W) placed concentrically (Fig. 9.2a).
This reactor was operated in a batch system with recirculation through a centrifugal pump (SIMES)
152 J. Femia et al.

Table 9.1. TOC and CP values for different rinsing.

1st rinsing 2nd rinsing 3rd rinsing Commercial product

TOC mg L−1 1502 28 19 488100


CP mg L−1 1127 21 15 366075

Table 9.2. Experimental conditions.

Variable Value

Chlorpyrifos initial concentration 15 mg L−1


TOC initial concentration 20 mg L−1
H2 O2 initial concentration 0–900 mg L−1
Input power lamp 20 W
Spectral fluence rate at the reactor 1.21 × 10−8 Einstein cm−2 s−1
λ
window, EP,0,W
Reaction time 4h
Total volume (VT ) 4L
Flow rate 3.5 L min−1
Initial pH 6 (natural)

and a feed tank. Figure 9.2b shows a schematic representation of the experimental device. Besides,
the experimental device had adaptations that allowed the sampling at the bottom of the tank.
In order to set the concentration of the emulsion to be treated, a triple rinsing was made on
an empty container of commercial chlorpyrifos. Each rinsing was made using a distilled water
volume of ¾ parts of the whole container volume. The water of each rinsing was analyzed by
TOC measurements. The initial concentrations of TOC and CP used in all the experiments were
chosen in accordance with those found in the 3rd rinsing (Table 9.1).
Finally, the emulsion to be treated was obtained by dilutions of the commercial concentrate of
chlorpyrifos in distilled water, adding the required amount of H2 O2 . The working emulsion was
stirred intensively before loading into the feed tank, which is really important when one works
with emulsions. During the course of each experiment, samples of 20 mL were taken at fixed
times for the tests in question.
λ
Measurements of the spectral fluence rate at the reactor window (EP,0,W ) were made with
potassium ferrioxalate actinometry (Murov et al., 1993). Table 9.2 summarizes the experimen-
tal conditions applied in the system. The molar ratio r is defined as r = CH0 2 O2 (mM)/CTOC 0
0
(mM), where CTOC is the total organic carbon initial concentration corresponding to chlorpyrifos
emulsion.
In this work, we studied the effect of varying the initial concentration of H2 O2 in the emulsion
degradation by keeping the remaining parameters constant (Table 9.2).

9.2.3 Analytical methods


Chlorpyrifos was analyzed by high pressure liquid chromatography (Waters 1525 Binary HPLC
Pump) with UV detector (Waters 2489 UV/Visible Detector), using an analytical column of C18
reverse phase (SUPELCOSILTM LC-18 – 4.6 × 250 mm). A solution of acetonitrile-water was
used as mobile phase (ratio 4:1), flow rate 1 mL min−1 , a temperature of 30◦ C and an injection
volume of 20 µL. The UV detector was set at 300 nm. The technique was adapted from the one
described in WHO/SIF/36.R3 (1999). Under this condition, the retention time for chlorpyrifos
was 4.5 min.
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 153

Hydrogen peroxide was analyzed using a colorimetric method (Allen et al., 1952), and
employing a Cary 100 Bio UV visible spectrophotometer (350 nm).
Total organic carbon (TOC) was analyzed with a Vario TOC Elementar. The pH was monitored
using an Orion Expandable ion Analyzer EA 940.

9.2.4 Bioassay test


For the Microtox test, a specially selected freeze-dried strain of the marine bacterium Vibrio
fischeri was used. The bioassays were performed using a Model 500 Analyzer (Strategic Diag-
nostics Inc.) according to the ASTM Standard Method D 5660-96 (2004). Hydrogen peroxide
present in the samples was removed prior to the toxicity analysis using catalase from bovine liver
after adjusting the sample pH between 6 and 8. The toxicity of the samples was tested with the
Microtox 81.9% screening test protocol (Boluda et al., 2002). The percentage of inhibition of
light emission was determined after 5 and 15 min of incubation.

9.3 RESULTS AND DISCUSSION

9.3.1 Preliminary runs


Two types of runs were made to investigate the eventual effects of UV radiation and H2 O2 ,
separately.
One run was made without UV exposure (not shown), starting from 15 mg L−1 chlorpyrifos and
681 mg L−1 of H2 O2 (r = 12). After 2 h of recirculation under normal operating conditions, no
degradation was observed. The other run was carried out without H2 O2 and under UV irradiation
for 2 h (r = 0). In this case, the photolysis of CP was observed; nevertheless, the conversion was
markedly lower than that obtained when H2 O2 was added (Fig. 9.3).

9.3.2 Effect of initial H2 O2 concentration


It is known that in the UV/H2 O2 process there is an optimum range of H2 O2 concentrations for
which the oxidation rate of the pollutant reaches maximum values (Mariani et al., 2010; Stefan
et al., 1996). It is not convenient to work with values of H2 O2 concentrations above the optimum
range because an inhibitory effect is produced in the presence of H2 O2 excess, in which hydrogen
peroxide captures hydroxyl radicals to form hydroperoxyl radicals. The hydroperoxyl radicals
have a lower oxidation capability; consequently, this reaction affects the contaminant degradation
rate. The values of the optimum H2 O2 concentrations depend on the concentration and type of the
pollutants in the effluent stream, i.e. on the rate constant of the reaction between the free radicals
and the pollutant and the rate constant of the recombination reaction. Because of the complexity
of the radical reaction, it is important to experimentally establish this operating optimum range
for each pollutant. Furthermore, the excess of H2 O2 is not desirable if the effluent is sent to
a biological treatment because H2 O2 is a powerful bactericide. Finally, but not less important,
working with a great excess of H2 O2 presents economic disadvantages.
0
Figure 9.3 shows the values of the relative concentrations of CP, CCP /CCP , versus reaction
0
time, analyzing different molar ratios between 0 and 16 (r). CCP is the initial concentration of
chlorpyrifos, working at the natural pH of the system, i.e., pH 6. As shown in Figure 9.3, when
the combination of UV plus H2 O2 is applied, the CP concentrations become negligible at 60 min
of reaction.
To better understand the effect of initial hydrogen peroxide on the kinetics of chlorpyrifos,
Figure 9.4 shows the conversion of chlorpyrifos against the initial hydrogen peroxide concentration
and r after 20 min of reaction. For H2 O2 concentrations higher than 390 mg L−1 , there is no marked
increase in conversion, assuming that initial molar ratios between 8 and 12 would be the optimal
value. Another important conclusion from this result is that it is not needed to work at r-values
higher than 8, because the excess of H2 O2 conspires against the economy of the whole process.
154 J. Femia et al.

Figure 9.3. 0 : 15 mg L−1 , H O initial


Relative concentration as a function of time for different r. CCP 2 2
concentrations: 0–900 mg L−1 .

Figure 9.4. Chlorpyrifos conversion percentage vs. r (or H2 O2 concentration) for a reaction time of 20 min.

There is a substantial difference between the conversions with r = 0 (without H2 O2 ) and those
with r-values > 0.
Conversions of 93% are reached in 20 min with r = 8 (Fig. 9.4), values highly advantageous
compared to those obtained with r = 0 (37% conversion). Even though it is difficult to perform
a strict comparison, it is worth mentioning that these conversion values were also reached for
similar times in the work of Murillo et al. (2010), in which other AOPs were applied to pure
chlorpyrifos in saturated aqueous solutions.
Figures 9.5a and 9.5b compares two experiments with ratios r = 2 and r = 12, presenting the
data of CP and H2 O2 concentration, and pH behavior as a function of reaction time. Despite the
great difference between the molar ratios, CP is reduced to a negligible amount before the first
hour in both cases. It can also be verified that CP concentration is reduced to less than 1 mg L−1
within the first 40 min of reaction. At r = 0, these values are reached only after 2 h.
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 155

Figure 9.5. Chlorpyrifos concentration, H2 O2 concentration and pH as a function of time for: (a) r = 2
(CH0 2 O2 = 106 mg L−1 ), (b) r = 12 (CH0 2 O2 = 485 mg L−1 ).

During the experiments, we worked at the natural pH of the preparations, which showed a rapid
initial decrease from 6 to 3.5 and then, with the progress of the oxidative degradation, a slight
increase in values (Figs. 9.5a and 9.5b). This behavior was observed in all experiments and could
be associated with the formation of organic acids followed by their transformation into the final
products of mineralization: inorganic acids and CO2 .
Figure 9.6a shows the temporal progression of the concentrations of the participating species
for the optimum operating condition r = 8. The observed loss of CP by the UV/H2 O2 process is
initially assumed to be pseudo first order with respect to CP concentration. If linearity is observed
when plotting ln = CCP /CCP 0
versus time (t), then this assumption would be valid. The slope of
regression line from such a plot yields the observed pseudo first-order rate coefficient kObs CP .
The same analysis can be done for H2 O2 . The plot of ln = C/C 0 versus reaction time shows a
straight-line behavior for both reactants (CP and H2 O2 ) (Fig. 9.6b). CP exhibited a pseudo first
order dependency with respect to itself. In H2 O2 , the same behavior was obtained. The observed
rate constants were kObs CP = 0.127 min−1 and kObs H2 O2 = 0.007 min−1 .
156 J. Femia et al.

Figure 9.6. 0 = 10.1 mg L−1 ;


(a) Chlorpyrifos and H2 O2 concentration evolution as a function of time, CCP
−1
CH2 O2 = 3391 mg L . (b) Pseudo first order decay curves of CP and H2 O2 .
0

9.3.3 Total organic carbon (TOC) evolution


In this work, the degree of mineralization of the samples was monitored through TOC mea-
surements, which yielded the results shown in Figure 9.7 at different r-values. The initial TOC
values are lower than the maximum TOC values determined in the course of the experiments. This
maximum was reached after around 20 min of treatment in each experiment, when the reaction
mixture changed from a murky emulsion to a clear one. This behavior could be explained by the
fact that CP has poor solubility and part of the emulsion adheres on the reactor walls. Conse-
quently, TOC-analyses are flawed with a systematic error, until all CP is dissolved. Alternatively,
the formation of the first soluble intermediates that still retain most of the organic carbon in their
chemical structures is a possible explanation of why the TOC values reach a maximum when the
mixture becomes clear (Fallman et al., 1999). Fallman et al. (1999) and Huston and Pignatello
(1999) observed similar results when they worked with other emulsified pesticides.
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 157

Figure 9.7. Temporal profiles of TOC and CCP at r = 6, 8 and 12.

The conversions reached were approximately 70% at 240 min in all runs (Fig. 9.7). The behavior
is independent of the initial concentration of H2 O2 in the range studied. In the case of r = 0, there
is no conversion (not shown); this implies that in the reaction time under study, the generated
intermediates would maintain the content of organic carbon.
During the degradation, the decline in TOC concentration was lower than the decline of CP
concentration (Fig. 9.7) for two reasons: the first one is that the mineralization is a rather long
sequence of oxidation reactions also implying acid/base equilibria. Therefore, the whole degrada-
tion kinetics must be similar or lower than the rate of disappearance of the initial component. The
second reason is that the emulsion contains other organic compounds that will be mineralized.
This was confirmed since the maximum TOC value of the emulsion (20 mg L−1 ) was higher than
the one assigned to the active ingredient (chlorpyrifos) alone (4.6 mg L−1 of TOC for 15 mg L−1
of CP, theoretical calculation). Furthermore, these organic compounds could affect to a greater or
lesser extent the rate of the pesticide degradation (Huston and Pignatello, 1999). Huston and Pig-
natello (1999) obtained similar results when they treated an emulsifiable concentrate of pesticide
(Lasso 4EC).

9.3.4 Evaluation of electrical energy per order


Since the photodegradation of aqueous organic pollutants is an electric-energy-intensive process,
and electric energy can represent a major fraction of the operating costs, simple figures of merit
based on electric energy consumption can be very useful and informative.
In the case of low pollutant concentrations (pseudo first order reactions), which applies here,
the appropriate figure of merit is the electrical energy per order (EEO ), defined as the number
of kW h of electrical energy required to reduce the concentration of a pollutant by 1 order of
magnitude (90%) in 1 m3 of contaminated water. The EEO (kW h m−3 ) can be calculated from the
following equations (Bolton et al., 2001):
P × t × 1000
EEO = (9.1)
V × 60 × log(Ci /Cf )
ln(Ci /Cf ) = k × t (9.2)

where P is the rated power (kW) of the AOP system, t the irradiation time (min), V the treated
wastewater volume (L), Ci and Cf the initial and final pollutant concentrations and k is the pseudo
158 J. Femia et al.

Figure 9.8. 0 =
Percentage of inhibition of Vibrio fischeri for a typical run. Experimental conditions: CCP
−1 −1
12.4 mg L ; CH2 O2 = 454 mg L ; pH 6; r = 10.
0

first order rate constant (min−1 ) for the decay of the pollutant concentration (Bolton et al., 2001).
From Equations (9.1) and (9.2), EEO can be written as follows:
3.84 × P
EEO = (9.3)
V ×k
For this work, the EEO calculated for CP removal and mineralization were 280 kW h m−3 and
889 kW h m−3 , respectively (under optimum operating conditions: 10.1 mg L−1 and 391 mg L−1
of CP and hydrogen peroxide, respectively, r = 8).
The standard figure of merit provides a direct link to the electrical efficiency (lower values
mean higher efficiency) of the AOP. It is interesting to compare these values with others obtained
for different pollutants using the UV/H2 O2 . For example, in Aleboyeh et al. (2008), the calculated
EEO values for C.I. Acid Orange 7 azo dye (AO7) color removal and mineralization were 1,133
and 5,691 kW h m−3 respectively (for CAO70
= 17.5 mg L−1 , CH0 2 O2 = 525 mg L−1 ).

9.3.5 Toxicity evaluation


The luminescence bacteria test with Vibrio fischeri, commonly called Microtox test, is a conve-
nient test to perform in a short time. The bioassay uses a suspension of Vibrio fischeri bacteria
and measures the reduction in light output of its natural luminescence on exposure to the toxic
compound of interest (Kaiser, 1998). Bacterial bioluminescence is related to cell respiration, and
any inhibition of cellular activity due to an interaction with a toxic substance results in a decreased
rate of respiration and a corresponding decrease of the rate of luminescence. Thus, luminescence
inhibition in the bacteria can effectively indicate toxic effects on higher organisms (Parvez et al.,
2006).
Figure 9.8 shows the percentage of inhibition of Vibrio fischeri for a typical run. At the begin-
ning, the toxicity was approximately 80% and, after 20 min of treatment, the percentage of
inhibition decreased to 60% and then increased again to 95%, possibly due to the generation of
more toxic intermediate products during the oxidation process.
The increased toxicity of the mixture could be explained by the possible formation of
chlorpyrifos-oxon (CPO), the most toxic metabolite of chlorpyrifos, or of other photoproducts
like 3,5,6-trichloro-2-pyridinol (TCP) (Kralj et al., 2007), or may be due to the interaction with
the other substances that exist in commercial formulations.
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 159

CP is classified as very toxic to invertebrate and aquatic organisms. In a study of the acute
toxicity of 11 organophosphates, CP was considered the most toxic insecticide for two marine
invertebrates (Artemia sp. and Brachionous plicatilis) (Guzella et al., 1997). Both CP and CPO
are potent cholinesterase inhibitors; cholinesterase is an enzyme vital for normal nerve function.
Tahara et al. (2005) found that the acute toxicity of CPO was greater than CP itself, due to its
active functional group (P=O).
In this study, the toxicity tends to a rapid decrease until 180 min of treatment, and then slowly
decreases up to approximately 10% inhibition. It should be remarked that even though more than
30% of the initial TOC content of the emulsion still remains, it is not necessary to complete the
mineralization because the intermediate organic compounds present after 240 min of treatment
show very low toxicity values (10% inhibition).

9.4 CONCLUSIONS

The intensive use of chlorpyrifos raises great concern because of the high solubility in water
and its toxicity and that of its first metabolites, which can decrease or rise along the natural
decomposition processes. This makes it necessary to develop effective, low-cost technologies for
the degradation of this compound and reduction of the impact of its residues.
The combination of hydrogen peroxide and UV radiation may be a suitable process to remove
chlorpyrifos from wastewaters. With the application of such process, a complete degradation of
chlorpyrifos was achieved at 40 min for initial H2 O2 concentrations between 390–490 mg L−1
(total volume treated 4 L, initial CP concentration 15 mg L−1 ). The grade of mineralization reached
in this work (70%) was good at reasonable times (240 min). The Microtox assay indicated that the
toxicity of the chlorpyrifos emulsion was significantly reduced before complete mineralization
and, consequently, it is not necessary to decompose completely the pollutant in order to obtain a
safe effluent.

ACKNOWLEDGMENTS

The authors are grateful to Universidad Nacional del Litoral (UNL), Consejo Nacional de Inves-
tigaciones Científicas y Técnicas (CONICET), and Agencia Nacional de Promoción Científica y
Tecnológica (ANPCyT) for financial support. Thanks are given to Elsa Grimaldi for the English
language editing.

REFERENCES

Abramovic, B.F., Banic, N.D. & Sojic, D.V.: Degradation of thiacloprid in aqueous solution by UV and
UV/H2 O2 treatments. Chemosphere 81 (2010), pp. 114–119.
Aleboyeh, A., Olya, M.E. & Aleboyeh, H. Electrical energy determination for an azo dye decolorization and
mineralization by UV/H2 O2 advanced oxidation process. Chem. Eng. J. 137 (2008), pp. 518–524.
Allen, A., Hochanadel, C., Ghormley, J. & Davis, T.: Decomposition of water and aqueous solutions under
mixed fast neutron and gamma radiation. J. Phys. Chem. 56 (1952), pp. 575–586.
ASTM Standard D5660-96: Standard test method for assessing the microbial detoxification of chemi-
cally contaminated water and soil using a toxicity test with a luminescent marine bacterium. West
Conshohocken, PA doi:10.1520/D5660-96R04, http://www.astm.org (accessed 2004).
Aydin, M.E., Ozcan, S. & Beduk, F.: Acute toxicity of organophosphorus pesticides and their degradation
by-products to Daphnia magna, Lepidium sativum and Vibrio fischeri. In: M. Stoytcheva (ed): Pesticides
– The impacts of pesticides exposure. ISBN 978-953-307-531-0, 2011, pp. 261–276.
Badawy, M.I., Ghalyb, M.Y. & Gad-Allaha, T.A.: Advanced oxidation processes for the removal of
organophosphorus pesticides from wastewater. Desalination 194 (2006), pp. 166–175.
Binimelis, R., Pengue, W. & Monterroso, I.: “Transgenic treadmill”: Responses to the emergence and spread
of glyphosate-resistant johnsongrass in Argentina. Geoforum 40 (2009), pp. 623–633.
160 J. Femia et al.

Bolton, J.R. Bircger, K.G. Tumas, W. Tolman, C.A. Figure-of Merit for the technical development and
application of advanced oxidation technologies for both electric-and solar-drived systems. Pure Appl.
Chem. 73 (2001), pp. 627–637.
Boluda, R., Quintanilla, J.F., Bonilla, J.A., Sáez, E. & Gamón, M.: Application of the microtox test and
pollution indices to the study of water toxicity in the Albufer Natural Park (Valencia, Spain). Chemosphere
46 (2002), pp. 355–369.
Briceño, G., Fuentes, M.S., Palma, G., Jorquera, M.A., Amoroso, M.J. & Diez, M.C.: Chlorpyrifos biodegra-
dation and 3,5,6-trichloro-2-pyridinol production by actinobacteria isolated from soil. Int. Biodeter.
Biodegr. 73 (2012), pp. 1–7.
Chioma Affam, A., Kutty, S.R. & Chaudhuri, M.: Solar photo-Fenton induced degradation of combined
chlorpyrifos, cypermethrin and chlorothalonil pesticides in aqueous solution. I. J. Chem. Environ. Eng. 6
(2012), pp. 167–173.
Fallmann, H., Krutzler, T., Bauer, R., Malato, S. & Blanco J.: Applicability of the photo-Fenton method for
treating water containing pesticides. Catal. Today 54 (1999), pp. 309–319.
García Molina V. Wet oxidation process for water pollution remediation. PhD Thesis, University of Barcelona,
Barcelona, Spain, 2006.
Giesy, J.P., Solomon, K.R., Coats, J.R., Dixon, K.R., Giddings, J.M. & Kenaga, E.E.: Chlorpyrifos: Eco-
logical risk assessment in North American aquatic environments. Rev. Environ. Contam. T. 160 (1999),
pp. 1–129.
Gomathi Devi, L., Narasimha Murthy, B. & Girish Kumar, S.: Photocatalytic activity of V5+ , Mo6+ and
Th4+ doped polycrystalline TiO2 for the degradation of chlorpyrifos under UV/solar light. J. Molec. Catal.
A: Chem. 308 (2009), pp. 174–181.
Guzella, L., Gronda, A. & Colombo, L.: Acute toxicity of organophosphorus insecticides to marine
invertebrates. Bul. Environ. Contam. T. 59 (1997), pp. 313–320.
Huston, P. & Pignatello, J.: Degradation of selected pesticide active ingredients and commercial formulations
in water by the Photo-assistted Fenton reaction. Water Res. 33 (1999), pp. 1238–1246.
Ikehata, K. & El-Din, M.: Aqueous pesticide degradation by hydrogen peroxide/ultraviolet irradiation and
Fenton-type advanced oxidation processes: A review. J. Environ. Eng. Sci. 5 (2006), pp. 81–135.
Kaiser, K.L.E.: Correlations of Vibrio fischeri bacteria test data with bioassay data for other organisms.
Environ. Health Persp. 106:2 (1998), pp. 583–591.
Knowles, A.: Recent developments of safer formulations of agrochemicals. Environmentalist 28 (2008),
pp. 35–44.
Kralj, M., Franco, M. & Trebse, P.: Photodegradation of organophosphorus insecticides – Investigations
of products and their toxicity using gas chromatography– mass spectrometry and AChE-thermal lens
spectrometric bioassay. Chemosphere 67 (2007), pp. 99–107.
Kundu, S., Chanda, A., Espinosa Marvan, L., Khetan, S. & Collins, T.: Facile destruction of formulated
chlorpyrifos through green oxidation catalysis. Catal. Sci. Tech. 2:6 (2012), pp. 1165–1172.
Manassero, A., Passalia, C., Negro, A.C., Cassano, A.E. & Zalazar, C.S.: Glyphosate degradation in water
employing the H2 O2 /UVC process. Water Res. 44 (2010), pp. 3875–3882.
Mariani, M., Labas, M., Brandi, R., Cassano, A. & Zalazar C.: Degradation of a mixture of pollutants in
water using the UV/H2 O2 process. Water Sci. Technol. 61 (2010), pp. 3026–3031.
Murillo, R., Sarasa, J., Lanao, M. & Ovelleiro, J.: Degradation of chlorpyriphos in water by advanced
oxidation processes. Water Sci. Technol. 10:1 (2010), pp. 1–6.
Murov S., Carmichael I. & Hug G.: Handbook of photochemistry. 2nd edition, Marcel Dekker, New York,
1993.
Parvez, S., Venkataraman, C. & Mukherji, S.: A review on advantages of implementing luminescence
inhibition test (Vibrio fischeri) for acute toxicity prediction of chemicals. Environ. Int. 32 (2006), pp.
265–268.
Pino, N. & Peñuela, G.: Simultaneous degradation of the pesticides methyl parathion and chlorpyri-
fos by an isolated bacterial consortium from a contaminated site. Int. Biodeter. Biodegr. 65 (2011),
pp. 827–831.
Resolución 456/09, Internet site: http://leg.msal.gov.ar/.
Sarkouhi, M., Shamsipur, M. & Hassan, J.: Metal ion promoted degradation mechanism of chlorpyrifos and
phoxim. Arab. J. Chem. (2012) DOI: 10.1016/j.arabjc.2012.04.026.
SENASA: Internet site: www.senasa.gov.ar.
Shemer, H. & Linden, K.G.: Degradation and by-product formation of diazinon in water during UV and
UV/H2 O2 treatment. J. Hazard Mater. 136 (2006), pp. 553–559.
Decontamination of commercial chlorpyrifos in water using the UV/H2 O2 process 161

Stefan, M., Hoy, A. & Bolton, J.: Kinetics and mechanism of the degradation and mineralization of acetone
in dilute aqueous solution sensitized by the UV photolysis of hydrogen peroxide. Environ. Sci. Technol.
30 (1996), pp. 2382–2390.
Tahara, M., Kubot, R., Nakazawa, H., Tokunaga, H. & Nishimura, T.: Use of cholinesterase activity as an
indicator for the effects of combinations of organophosphorus pesticides in water from environmental
sources. Water Res. 39 (2005), pp. 5112–5118.
WHO/SIT/21.R3 Technical Chlorpyrifos: Full specification. Internet site: https://apps.who.int/ctd/docs/
whopes/new_docs/full_spec/chlorpyrifos.pdf (accessed 1999).
Wu, C. & Linden, K.: Phototransformation of selected organophosphorus pesticides: Roles of hydroxyl and
carbonate radicals. Water Res. 44 (2010), pp. 3585–3594.
Zhang, Y., Hou, Y., Chen, F., Xiao, Z., Zhang, J. & Hu, X.: The degradation of chlorpyrifos and diazinon in
aqueous solution by ultrasonic irradiation: Effect of parameters and degradation pathway. Chemosphere
82 (2011), pp. 1109–1115.
This page intentionally left blank
CHAPTER 10

Abatement of nitrate in drinking water. A comparative study of


photocatalytic and conventional catalytic technologies

F. Albana Marchesini, Guadalupe Ortiz de la Plata, Orlando Alfano,


M. Alicia Ulla, Eduardo Miró and Alberto Cassano

10.1 INTRODUCTION

The contamination of water due to both intensive fertilization and waste effluents from industries
has produced an increase of nitrate concentration in groundwater. Nitrates are reduced to nitrites
in the digestive system, affecting hemoglobin and impairing its function as oxygen-carrier, thus
causing the “blue baby syndrome”. They are also related to several forms of cancer, e.g. ovarian
and prostate cancers. This contamination problem has led to the introduction of guidelines to
establish upper limits for the concentration of nitrates and related species (nitrites and ammonia)
in water for domestic use. The Drinking Water Directive- European Commission (98/83/EC)
sets a maximum allowable concentration of nitrate of 50 ppm, 0.1 ppm for nitrite and 0.5 ppm
for ammonia. The World Health Organization recommends a maximum nitrate concentration of
10 mg L−1 in drinking water. The U.S. Environmental Protection Agency (EPA) has established a
maximum contaminant level (MCL) in drinking water of 10 mg L−1 as nitrate–N to protect infants
from methemoglobinemia (Ward et al., 2005).
At present, the most widespread technologies for the removal of nitrates are biological denitrifi-
cation and physico-chemical processes, namely ion exchange, reverse osmosis and electro dialysis.
However, they present serious problems, i.e. bacterial processes include handling difficulties, low
reaction rates, need for removal of by-products and low space velocities. Physico-chemical pro-
cesses usually transform the nitrate into a brine that has to be treated or disposed afterwards.
Moreover, some of these technologies can be expensive. Therefore, increasing attention has lately
been paid to a novel technology, still in its development stages, catalytic denitrification (Ilinitch
et al., 2000; Pintar et al., 1999), which employs solid bimetallic catalysts. In this catalytic process,
nitrates are reduced to nitrogen using hydrogen (Reaction 10.1); however, undesirable products
such as nitrite and ammonium are also formed (Reactions 10.2 and 10.3).
2NO− −
3 + 5 H2 → N2 + 2OH + 4 H2 O (10.1)

NO− −
3 + H2 → NO2 + H2 O (10.2)
2NO−
3 +8 H2 → 2NH4+ −
+ 4 HO + 2 H2 O (10.3)
It is accepted that the reaction mechanism involves the reduction of nitrate into nitrite, which
is in turn reduced towards nitrogen or over-reduced to give ammonia (Pintar et al., 1999). Thus,
the positive catalytic effect is related to the ability of the catalyst to selectively hydrogenate nitrite
into nitrogen. In general, the most active catalysts for nitrate reduction show the lowest nitrite
formation, because they are also active for nitrite reduction (Marchesini et al., 2008).
As shown by reactions 1.1 and 1.3, nitrate reduction produces OH− ions and their local accu-
mulation could negatively affect the catalytic activity and the selectivity to N2 . Therefore, a pH
of ca. 5 was maintained during the reaction time by the addition of controlled amounts of HCl
(Garrón et al., 2005) in order to improve the selectivity to N2 .
163
164 F.A. Marchesini et al.

Several solid catalysts have been proposed by various authors (Deganello et al., 2000; Matatov
et al., 2000; Strukul et al., 2000). The catalysts were prepared over different mesoporous supports
such as massive oxides, alumina or silica, with the addition of noble metals such as Pt or Pd as the
main metal, and a second metal such as Cu, Co or In as promoter metal. The possible mechanism
for the catalytic reduction is through the combination of active sites in the bimetallic catalyst
(Prusse et al., 2000). The nitrate is reduced to nitrite and the later is reduced by the noble metal
to nitrogen or ammonium depending on the site selectivity and the environmental conditions, as
mentioned above.
The restriction about a catalytic process for nitrate removal from drinking water is severe, i.e.
the product stream must conform to drinking water standards and the energy consumption must
be low for the process to be cost effective; therefore, temperature and pH cannot be adjusted
freely, and no toxic substances may leach from the catalyst (Bems et al., 1999).
The reactions involved in photocatalysis are mostly oxidation reactions, which utilize the pos-
itive holes produced by illuminating the semi-conductor, i.e. TiO2 , to generate oxidizing radicals
by reaction with adsorbed HO− or water. However, this particular application, nitrate abatement,
calls for a reductive reaction. The irradiation of titanium dioxide with the appropriate wavelength
gives rise to the formation of not only positive holes (vacancies) but also photoelectrons. Thus,
this less widely studied reaction path might provide the necessary conditions to achieve nitrate
reduction that is the ultimate target of this work.
The photocatalytic route is represented by:
UV + MO → MO(h+ + e− ) (10.4)

Here, MO stands for metal oxide, usually a semiconductor such as TiO2 . The employed UV
must be of the required wavelength in order to overcome the semiconductor band gap.
The oxidative path is usually represented by:
h+ + H2 O → H+ + HO• (10.5)

while the reductive reaction is mostly presented as:


e − + O 2 → • O−
2 (10.6)

However, in the absence of oxygen, another reductive reaction may occur. In the presence of
an appropiate catalyst electrons produce hydrogen:
2e− + (Cat) + 2H+ → H2 (10.7)

The photoreduction of nitrates was studied by Ranjit et al. (1997); they proposed that metals
with high affinity for electrons need to be added to TiO2 . Once the electrons are trapped in this
way, they can be used as reducing agents for nitrates. Unfortunately, as opposed to what happens
in oxidation practices, the reactions employed in reduction processes have not been so extensively
studied.
Nevertheless, the possibility of using various forms of titanium dioxide modified by metals
with the ability to capture electrons has acquired new impetus and increasing importance. The
photocatalytic reduction of nitrate to nitrogen in aqueous solution was investigated in presence
of formic acid as a hole scavenger to improve the photocatalytic efficiency, using different doped
catalysts: Ag-TiO2 (Zhang et al., 2005) and Bi+ 3 -TiO2 (Rengaraj and Li, 2007). This possibility
has been also explored in studies of reductive reactions (Sakthivel et al., 2004; Shiraishi et al.,
2004; Zhang et al., 2006; Cohen, 2001), hydrogen production from water (Chiarmello et al.,
2011; Nada et al., 2008; Kudo, 2003), and the reduction of carbon dioxide emissions to the
atmosphere (Stock et al., 2011).
In this work, Pd-In metallic couples have been selected on the grounds that In promoted
catalysts have good activity for nitrate conversion and a potential high selectivity to N2 , as reported
elsewhere (Marchesini et al., 2012). TiO2 has been selected as support because it can perform as
electron supplier in the presence of UV radiation.
Abatement of nitrate in drinking water 165

The aim of this preliminary work is to compare the efficiency of a catalyst having a titanium
dioxide support and Pd-In as metallic additives operating under conventional reaction conditions
with that of the same catalyst but irradiated with UV radiation. The purpose of such a comparison is
to determine if its photocatalytic activity allows a better control of the reductive function and avoids
the formation of undesirable reaction products, specifically nitrites and NH+ 4 . If the results of this
exploratory study were attractive enough, possible follow-ups would include the optimization of
this catalyst composition and kinetic studies to characterize all the reaction parameters.

10.2 MATERIALS AND METHODS

10.2.1 Chemicals
The following reagents were used: Aqueous solutions of PdCl2 (10 mg mL−1 ) dissolved in con-
centrated HCl and aqueous solutions of In2 O3 (4.6 mg mL−1 ) dissolved in concentrated HCl, TiO2
(Degussa, P25) was used as support and deionized water was employed in all experiments.

10.2.2 Catalyst preparation


The TiO2 powder was suspended in an aqueous solution of Pd chloride and In chloride in order
to obtain concentrations of 1 wt% Pd and 0.25 wt% In, respectively. The water was evaporated
and the solid was dried overnight at 120◦ C and calcined at 500◦ C in air flow for 4 h. Before the
hydrogenation reaction, the catalysts were reduced under H2 flow at 450◦ C for 1 h in order to
reduce the metals. The catalyst so obtained was used in all catalytic experiments without additional
pre-treatment.

10.2.3 Catalyst characterization


10.2.3.1 X-Ray diffraction analysis (DRX)
X-ray diffractometer patterns were acquired with an XD-D1 Shimadzu instrument, using CuKα
radiation at 30 kV and 40 mA. The scan rate was 1◦ min−1 in the 2θ = 10–80◦ range.

10.2.3.2 Temperature-programmed reduction (TPR)


Temperature-Programmed Reduction (TPR) experiments were performed with 50 mg of catalyst
placed in a quartz flow reactor using an Okhura TS 2002 S instrument equipped with a TCD
detector. The reducing gas was 5% H2 in Ar, flowing at a rate of 30 mL min−1 , and the heating
rate was 10 C min−1 .

10.2.4 Catalytic activity measurements


10.2.4.1 Preliminary batch experiments
The reaction test was performed under batch operation in a reactor (250 mL) equipped with a
magnetic stirrer (Figure 10.1). Experiments performed at different stirring rates between 500 and
1000 rpm showed that the reaction was limited by mass-transfer below 600 rpm. Between 600
and 1000 rpm, external diffusion limitations were overcome and similar activities were observed.
Consequently, a rate of 800 rpm was chosen for all the final experiments. The hydrogen flow was
set over 400 mL min−1 in order to overcome the limitations in availability of dissolved hydrogen.
The pH value was measured using an automatic pH controller unit. A gas bubbling device, which
allows maintaining a flow rate of 400 mL min−1 , was used to bubble either N2 or H2 . The reaction
tests were carried out at room temperature, pH = 5 by adding drops of HCl 0.1 M and atmospheric
pressure.
The stirred batch reactor was loaded with 80.0 mL of distilled water and purged with N2 flow
for 20 min. Then, 200 mg of the reduced Pd,In/TiO2 catalyst was added and the N2 purge was
166 F.A. Marchesini et al.

Figure 10.1. Schematic representation of the batch experimental setup. (1) Reactor; (2) H2 or N2 flow inlet;
(3) magnetic stirrer; (4) Sample port; (5) HCl injection and pH meter; (6) N2 supply; (7) H2
supply.

repeated. After that, a hydrogen flow was fed to the reactor and the reaction started when a certain
amount of a nitrate solution was added to the reactor to obtain 100 N-ppm of nitrate as initial
concentration.
Before the reaction experiments, the catalysts were pre-treated under a H2 flow of 100 mL
min−1 at 450◦ C with a heating rate of 10◦ C min−1 in order to obtain the reduced metallic forms
of the Pd and In species on the surface.

10.2.4.2 Photocatalytic experiments


Photocatalytic experiments were performed in a more complex reacting system. Figure 10.2 shows
a schematic representation of the experimental setup employed in the case when the catalytic
irradiated degradation of nitrates was performed. The reaction was carried out in a cylindrical
reactor made of stainless steel, with a lining wall made of Teflon. It was operated as a slurry
reactor inside the loop of a batch recycling system. The reactor had two circular, flat windows
made of borosilicate glass. The employed photoreactor total volume was 29 mL. Radiant energy
was supplied by two sets of four UV lamps (TL 4W/08 black light UVA lamps from Phillips),
placed in front of each window. The lamps emitted in the 300–400 nm wavelength range with a
principal emission peak at ca. 350 nm. Ground glass plates, situated between the lamps and the
reactor windows, were used to produce diffuse inlet radiation. This modification is an important
addition to the reactor, in order to facilitate the description of the existing radiation field, if a
rigorous kinetic interpretation of the reaction kinetics is needed. The storing tank was equipped
with a sampling valve, a pH measurement device, an HCl injection port, a gas inlet for hydrogen
supply and completed with a water circulating jacket to secure isothermal operating conditions.
The total liquid system volume was 400 mL, rendering a ratio of:
VR,irr /VTot. = 0.0725 VR,irr volume under radiation; VTot. total volume

The system was maintained under hydrogen overpressure to guarantee a reductive reaction
media during the photochemical reaction. A good mixing operation was provided during the
reaction time. This was achieved by resorting to a pump with a flow rate regulator that adjusted
the recirculation performance, thus guaranteeing low conversion per pass in the reactor and
uniform concentration of the catalyst throughout the system. Each run lasted 2 h and samples
Abatement of nitrate in drinking water 167

Figure 10.2. Schematic representation of the experimental setup. [Adapted from Satuf et al. (2007)] (1)
Reactor; (2) Windows; (3) Tubular lamps; (4) Tank; (5) Sample port; (6) Thermostatic bath;
(7) H2 supply; (8) Pump, (9) pH meter and (10) HCl injection. When the reacting system was
used without irradiation, the lamp power inputs were turned off.

Table 10.1. Experimental conditions (*).

Parameter Value

Reactor inner diameter 8.6 cm


Reactor length 0.5 cm
Reactor volume 29 mL
Total system volume 400 mL
Lamp nominal power (four on each side) 4W
Lamp emission range 300–400 nm
Thermostatic bath temperature 25◦ C
Nitrates initial concentration 100 N ppm
Reaction time 2h
pH 5

(*) The reactor volume indicated: 29 mL corresponds to the irradiated


section of the total system volume.

were taken from the tank at fixed time intervals. The dimensions and main characteristics of the
experimental setup are summarized in Table 10.1.
Before each run, the water suspension of the catalyst was treated according to the following
protocol: (i) sonication in order to ensure uniform dispersion of the catalyst particles, (ii) N2
bubbling and (iii) Hydrogen bubbling for a period of 20 minutes. To start the reaction, a calculated
volume of a concentrated solution of KNO3 was added to obtain an initial concentration of 100 N
ppm NO− 3 . To initiate the irradiated runs, the power of both lamp systems was simultaneously
turned on. Notice that, even if using the same equipment as the one described in Figure 10.2,
differing from the case when no radiation was present, hydrogen bubbling was only used during
the first 30 minutes of the photocatalytic experimental run that lasted a total of 120 minutes.
Reproducibility tests were performed for this run.
The gas phase was not analyzed. The residual concentration of nitrate, the production of nitrite
and ammonium were analyzed in samples taken from the tank. Then, when the conversion was
168 F.A. Marchesini et al.

40%, the nitrogen selectivity (SN2 ) was calculated from the mass balance of nitrogen products
(nitrite and ammonia analyzed in the solution) according to:
SN2 = 100 − SNO2 − SNH4 SN2 = selectivity to N2 ; SNO2 = selectivity to NO2 ;
SNH4 = selectivity to NH4+

10.2.5 Analytical methods


In order to determine nitrates, nitrites and ammonium concentrations, small samples were taken
from the vessel and analyzed using Vis spectroscopy (Cole Parmer 1100 Spectrophotometer)
combined with colorimetric reagents. The cadmium column method and the colorimetric Gries
reaction were used (APHA, Standard methods for examination of water and wastewater, 1992).
This colorimetric reaction is the same as the one employed in the assay for nitrites. Ammonium was
analyzed by the adapted Berthelot method (Grasshoff et al., 1983). The other nitrogen containing
compounds were N2 and small amounts of N2 O or NO in the gas phase, which were not measured.

10.3 RESULTS AND DISCUSSION

10.3.1 Physicochemical characterization of the Pt,In/TiO2 catalyst


The optical characterization of the catalyst is important from two points of view: (i) it is essential
for the correct design of the experimental reactor avoiding the appearance of dark volumes that
do not contribute to the reaction and (ii) to support the discussion of the observed favorable
differences that offers the photocatalytic reaction in removing nitrates without the formation of
undesirable by-products.
A single beam, high precision spectrophotometer was used for measuring the catalyst optical
properties (Optronic OL Series 750 spectroradiometer equipped with an OL 740-70 integrating
sphere reflectance attachment). Suspensions were prepared having the same pH as in the experi-
ments. Immediately after, they were sonicated for 20 minutes. For reflectance measurements, the
cell was calibrated against a PTFE accessory that reflects 99%+ of the reaching radiation. The
standard sampling cell of the instrument is 1 cm. More details concerning these experiments can
be found in Satuf et al. (2005) and Ortiz de la Plata et al. (2008).
With respect to absorption, adopting water as a reference, the observed transmittance was one.
Therefore, new measurements, also made at pH = 5 were carried out but resorting to different
degrees of dilution. The data are shown in Figure 10.3. As a result of the information provided
by this figure, it can be concluded that for the reactor used in this work, which has a thickness of
0.5 cm, it is not be desirable to use concentrations of catalyst much larger than 1 g L−1 because
it would be very close to the limit of functioning having parts of its volume totally opaque and
useless for the photocatalytic reaction.
Using the integrating sphere accessory of the Optronic spectrophotometer, according to the
details described in Ortiz de la Plata et al. (2008) diffuse reflectance measurements were made. For
these measurements solid pellets of the catalyst were built. Differing from previous experiments,
since the catalyst pellet was completely opaque, there was no need to use a black body (a light
trap) on the back of the pellet.
Note that none of these data corresponds to the employed concentration inside the reactor. The
data for the catalyst loadings in the reactor (1 g L−1 ) resulted in transmittance values equal to 0
for all wavelengths. However, these measurements were made within the range of the accuracy of
the instrument, to have complete information about the characteristics of the catalyst employed.
For a given surface, the diffuse reflectance is defined for each wavelength, as the ratio between
the incident and the reflected radiation flux. It is obtained by comparison with a standard sample
in two separate experiments:
Rλ = q− + −
λ /qλ R = diffuse reflectance; qλ = reflected radiation flux;

q+
λ = incident radiation flux.
Abatement of nitrate in drinking water 169

Figure 10.3. Transmittance measurements. Concentrations of the catalysts are expressed in g/L.

Figure 10.4 shows the ability of the catalyst to reflect radiation. However, as compared with the
original support (Degussa P 25), its value is considerably reduced. The information exhibited in
this figure, analyzed in conjunction with those of the previous one, is of primary importance. They
show that the modified catalyst used in these experiments has a reflectance significantly lower
than the typical standard represented by a variety of titanium dioxide provided by Degussa. If the
reflectance between approximately 320 and 400 nm has diminished, it provides an indirect, but
equally valid confirmation that the radiation absorption of the catalyst in a very considerable part
of the usable wavelengths has not diminished; on the contrary, the figure even allows inferring
that it would have been increased.
This information is very useful for the analysis of the reaction behavior in the photocatalytic
reactor because it deals with a photoreactor inside a recycle with a rather large storage tank,
where the irradiated volume/dark volume ratio is equal to 0.078. In this regard, it is important
to emphasize that in a photocatalytic reactor the optical properties and, more specifically, the
absorption of radiation in the appropriate range of wavelengths, has the same importance as the
specific surface area in the conventional catalytic reactions and, in some cases, even more.
The three XRD patterns shown in Figure 10.5 clearly indicate that the crystalline frameworks
of Pd/TiO2 and Pid,In/TiO2 are practically similar to those of bare TiO2 (P25), which is composed
of ca. 80% anatase and ca. 20% rutile phases (Matos et al., 2012).
170 F.A. Marchesini et al.

Figure 10.4. Reflectance measurements.

Figure 10.5. DRX of: (A) Bimetallic catalyst Pd,In 1:0.25 (B) monometallic Pd 1%wt catalyst (C) TiO2 .
Abatement of nitrate in drinking water 171

Figure 10.6. Temperature-programmed reduction of: (A) fresh Pd,In/TiO2 ; (B) Used (H2 - Pd,In/TiO2 ).
Experimental conditions see 2.3.2.

The TPR results confirm the presence of the reducible species on the catalytic surface for
the bimetallic catalyst after calcinations at 400◦ C (fresh Pd,In/TiO2 , Figure 10.6. (A)). The H2
consumption at 480◦ C is related to the reduction of TiO2 surface (Yang et al., 2012); at 330◦ C, it
corresponds to the reduction of interacting Pd-In particles. At lower temperatures, a negative peak
is identified, which is due to the decomposition of β-palladium hydride. This species is easily
formed when small Pd particles are in contact with H2 at low temperature (Sá et al., 2006).
The TPR profile of the catalyst after reduction in H2 flow at 450◦ C and being used under
reaction conditions indicates that the surface species remain reduced (Figure 10.7. (B)). However,
the negative signal associated with the decomposition of β-palladium hydride is more intense than
that of the fresh catalyst, suggesting that some agglomeration of the metallic particles takes place
during the catalytic reaction.

10.3.2 Catalytic reduction of nitrates: Conventional batch reactor


Figure 10.7 shows the evolution of NO− − +
3 , NO2 and NH4 during the reaction when the Pd,In/TiO2
was used as catalyst and tested in the conventional batch reactor system.
The catalyst presented high activity for nitrate reduction under hydrogen as reductant, simi-
lar to that presented in previous studies for the Pd,In/Al2 O3 catalyst (Marchesini et al., 2008).
Nitrate reached a conversion of 100% in less than 45 min. Besides, the nitrite concentration had
a maximum around 20 min and dropped almost to zero after 50 min. This behavior is in line with
its participation as an intermediate species to produce either N2 or NH+ 4 . Moreover, the nitrite
over-reduction to give ammonia was significant obtaining a S+ NH4 above 20%. Consequently, the
Pd,In/TiO2 formulation presents a good catalytic performance for nitrate reduction. However,
under these conditions, the formation of a significant concentration of an undesired byproduct
(NH+ 4 ) was observed.
This is a clear indication of the need to search for a reaction system that can enable better
possibilities of controlling the extension of its reducing activity. A controlled indirect reduction of
nitrates seems possible by irradiating TiO2 with appropriate UV radiation,whereby photoelectrons
could be generated and used as reductant agents of NO− x (x = 2 or 3).
172 F.A. Marchesini et al.

Figure 10.7. Evolution of nitrate, nitrite and ammonia as a function of time during the nitrate reduction
under hydrogen in the system shown in Figure 10.1. Reduction of nitrates in the conventional
reactor.

10.3.3 Catalytic reduction of nitrates: Photocatalytic reactor


Figure 10.8 shows the composition profiles of NO− − +
3 , NO2 and NH4 for the Pd,In/TiO2 catalyst
evaluated employing the experimental arrangement shown in Figure 10.2, without UV light. In
general, it can be seen that during the reaction time, the Pd,In/TiO2 catalyst reduces nitrates up
to 40% and the main product is NO− +
2 with negligible formation of NH4 . As compared with the
previous results in the batch reactor (Figure 10.7), while the nitrate conversions are lower in
Figure 10.8, a higher selectivity to nitrite can be noticed, but the selectivity to ammonia is slightly
lower. These results indicate that the fluid dynamic conditions of the experimental reactor shown
in Figure 10.2 strongly pose a limit to nitrate conversion and affect selectivities. A summary of
the data is depicted in Table 10.2, where the performance of the Pd,In/Al2 O3 catalyst in the batch
reactor is also shown for comparison purposes.
In any event, these experiments are unquestionably necessary to be able to compare them
with those carried out in the photocatalytic reaction, so as to contrast the results obtained under
identical operating conditions.
Figure 10.9 shows the results obtained employing the photocatalytic reactor, i.e. the same
Pd,In/TiO2 catalyst but subjected to UV irradiation. Under these conditions, electron and vacan-
cies are generated and the addition of metal clusters improves the numbers of electrons reaching
the surface for the reaction to take place (Sá et al., 2008; Anpo et al., 2003; Jakob et al., 2003).
The most important observation is that the reaction with UV light improves the selectivity to N2
and decreases the selectivity to nitrite.
Despite the lower nitrate conversions as compared with the batch reactor, the results obtained
are positive because ammonia is a non-desirable by-product caused by nitrite over-reduction.
Thus, a catalytic system with a low selectivity to this product and a high selectivity to nitrogen is
highly desirable.
Abatement of nitrate in drinking water 173

Figure 10.8. Composition profile as a function of time during the nitrate reduction under hydrogen in a
reacting system shown in Figure 10.2 without UV radiation.

Figure 10.9. Composition profile as a function of time during the nitrate reduction under hydrogen in the
reacting system shown in Figure 10.2 employing UV radiation.

The experimentally observed nitrate conversions were equal (cf. Table 10.2 40%), operating
either with or without UV irradiation. The outstanding difference yielded by the photocatalytic
process was that the conversion rendered a 98.1% selectivity with respect to N2, while under
dark conditions; the same variable was only 67.5%. As stated above, NH4+ was not detectable.
This interesting positive effect could be ascribed to the improvement in both the redox behavior
and adsorption properties of the catalyst when subjected to UV radiation. Thus, both redox and
174 F.A. Marchesini et al.

Table 10.2. Results of percent conversion (%X) at 120 min of reaction time. No irradiation. Selectivities to
nitrogen (SN2 ); to nitrite (SNO2 ) and ammonia (SNH4 ) at X = 40% conversion.

Catalysts Operating conditions %X SNO2 SNH4 SN2

Pd,In/Al2 O3 Conventional/H2 /Batch reactor 100 10.0 11.5 78.5


Pd,In/TiO2 Conventional/H2 /Batch reactor 100 20.0 10.0 70.0
Pd,In/TiO2 Photocatalytic reactor (Ph-R) 40 32.5 Nda 67.5
Radiation power off. /H2 )
Pd,In/TiO2 H2 /Photocatalytic reactor (Ph-R) 40 1.9 Nd 98.1
a Nd: Non detectable

Figure 10.10. Local Volumetric Rate of Photon Absorption for the photocatalytic reactor and different cat-
alyst loadings. (——): Cm = 1.0 g L−1 ; (– – –): Cm = 0.5 g/L; (· · ·): Cm = 0.1 g/L; (– · – ·):
Cm = 0.05 g/L. Reprinted from Applied Catalysis B: Environmental, 82, M.L. Satuf, R.J.
Brandi, A.E. Cassano, O.M. Alfano, Photocatalytic Degradation of 4-Chlorophenol: A
Kinetic Study, 37–49, Copyright (2008), with permission from Elsevier.

adsorption properties are key steps and determine the selectivity towards ammonia and nitrogen.
In this work, we report the positive effect of UV radiation. Further research work is under way
for a better understanding of this novel observation.

10.3.4 Spatial Distribution of the Radiation Absorption


Since the catalyst support has been optically characterized in previous works (Satuf et al., 2005),
it is possible to predict the spatial distribution of the availability of radiation for the formation of
holes and electrons inside the reactor, i.e. it is possible to calculate the Local Volumetric Rate of
Photon Absorption (LVRPA) for different working catalyst loadings.
Figure 10.10 shows the results of these calculations for the same photocatalytic reactor, with
a reactor length of 0.5 cm, using a different titanium dioxide catalyst and four catalyst loadings
(1.0, 0.5, 0.1, and 0.05 g L−1 ). The solid line corresponds to 1.0 g L−1 , equivalent to the catalyst
load used in the experiments, and the dotted line corresponds to a load of 0.1 g L−1 , the maximum
concentration that was possible to use in the spectrophotometer measurements without dilution.
As shown in the figure, for the lower catalyst loadings of 0.05 and 0.1 g L−1 , the spatial variations
of the radiation field is insignificant. Additionally, it can be noticed that even for the highest
catalyst loading of 1.0 g L−1 , the photoreactor is fully irradiated.
Abatement of nitrate in drinking water 175

As a good approximation, it is possible to consider that the used Pd,In/TiO2 catalyst loading of
1.0 g/L would give LVRPA values almost equivalent to those shown in Figure 10.10. On the basis
of this assumption, one might conclude that a photocatalytic reactor having a thickness of 0.5 cm
would also be almost fully irradiated. From the photocatalytic point of view, by employing these
low optical thicknesses, the reactor can operate under much better conditions because of three
reasons: (i) the photon absorption will be surely equally achieved in the whole reaction volume,
(ii) the external mass transfer control produced by catalyst agglomeration will be minimized, and
(iii) the photonic effectiveness will be maximized (Ballari et al., 2008). Moreover, a much larger
volume can be undoubtedly fully illuminated, resulting in a much better relationship of Irradiated
Volume/Total Volume. This new operation condition can lead to very important consequences as
explained below.
For photocatalytic oxidations in recirculating systems, it has been shown that the formation
of holes and electrons that participate in degradation reactions takes place only in the irradiated
volume (Cassano and Alfano, 2000). This renders a mass balance as shown below:

dCi (x, t)  VR,irr
= Ri (x, t)VR,irr
dt VTank VTotal

where the symbol   indicates the volume averaged value of the reaction rate Ri over the
volume VR,irr and VTank of the tank volume. This equation shows that the contribution of the
photocatalytic reaction is affected by the ratio of VR,irr /VTotal .
The reaction system studied in this case, may be conceived, globally, as a reactor that has two
parts and where two different types of reactions occur in parallel: (i) a set of chemical changes that
may occur in the entire volume (dark and irradiated reaction spaces) and (ii) a group of physical
and chemical phenomena that may only occur in those parts where there are photons (irradiated
part). In these cases, the mass balance is unfolded according to the following equation (Rossetti
et al., 2002):
    
dCi  VR,irr VT − VR,irr
= Ri,Irr+Dark (x, t)VR,irr + Ri,Dark (x, t)VR,irr
dt VTank VTotal P+T VTotal T

where P = photo, T = thermal.


When comparing the behaviors of the irradiated and non-irradiated systems, one might concieve
that the nitrite observed in the photocatalytic process corresponds to the reactions that occur in the
large proportion of existing dark volumes. As a result, according to the analysis that can be made
in the previous equation, the selectivity for N2 inside the irradiated section could reach a value
of almost 100%. Moreover, it is quite probable that even a portion of the nitrite generated on the
dark part, may have been degraded in the illuminated area. This may indicate that in a practical
operation, an optical thinner reactor (avoiding opaque volumes) and an almost fully irradiated
system (instead of the recirculating system having a very significant dark volume) should have a
better performance.

10.4 CONCLUSIONS

The conversion achieved in the irradiated system after 120 minutes of reaction (cf. Table 10.2
40%) is equivalent to that obtained with the same device operated in the dark.
Under these conditions, the photocatalytic system presents selectivities for N2 which are sub-
stantially better. These results represent a significant progress in the process. The irradiated system
achieved a 98.1% selectivity with respect to N2 while the non-irradiated one reached 67.5%. In
addition, the dark process has a serious drawback, i.e., it renders 32.5% selectivity with respect
to nitrite, which constitutes an undesired outcome.
176 F.A. Marchesini et al.

These promising results pave the way towards the optimization of the use of photocatalytic
systems to reduce nitrates. The combination of radiation and heterogeneous catalysis may help in
tuning the redox and adsorption properties that are the key steps in the reaction mechanism.
It should also be noted that the reported experiments used a metal loading that is four times
lower than the ones previously reported and a lower ratio of catalyst loading with respect to the
concentration of contaminant, which constitutes a substantial economic improvement.

ACKNOWLEDGMENTS

The authors are grateful to Universidad Nacional del Litoral (UNL), Consejo Nacional de Inves-
tigaciones Científicas y Técnicas (CONICET) and Agencia Nacional de Promoción Científica
y Tecnológica (ANPCyT) for their financial support. Thanks are given to Elsa Grimaldi for the
English language editing.

REFERENCES

APHA. American Public Health Association. Standard methods for the examination of water and wastewater.
Washington, DC, (1992) 312 pp.
Anpo M., Aikawa N. and Kubokawa Y., The design and development of highly reactive titanium oxide
photocatalysts operating under visible light irradiation, J. Phys. Chem. 88 (1994) 3998.
Ballari M., Brandi R., Alfano O. and Cassano A., Mass transfer limitations in photocatalytic reactors employ-
ing titanium dioxide suspensions: II. External and internal particle constrains for the reaction, Chemical
Engineering Journal 136 (2008) 242.
Bems B., Jentoft F. and Schlögl R., Photoinduced decomposition of nitrate in drinking water in the presence
of titania and humic acids, Appl. Catal. B: Environm. 20 (1999) 155.
Cassano A. and Alfano O., Reaction engineering of suspended solid heterogeneous photocatalytic reactors,
Catalysis Today 58 (2000) 167.
Chiarmello G., De Paola A., Palmisano L. and Selli E., Effect of titanium dioxide crystalline structure on the
photocatalytic production of hydrogen, Photochem. Photobiol. Sci. 10 (2011) 355.
Cohen D. and Ray A., Removal of toxic metal from wastewater by semiconductor photocatalysis, Hem. Eng.
Sci. 56 (2001) 1561.
Daum J. and Vorlop K., Kinetische undersuchung der katalytitischen nitratreduktion: Aufbau des
teststandes, Chem. Eng. Tech. 70 (1998) 1567.
Deganello F., Liotta L., Macaluso A., Venezia A. and Deganello G., Catalytic reduction of nitrates and nitrites
in water solution on pumice-supported Pd–Cu catalysts, Appl. Catal. B: Environmental 24 (2000) 265.
Garron A., Lázár K. and Epron F., Use of formic acid as reducing agent for application in catalytic reduction
of nitrate in water, Appl. Catal. B: Environmental 59 (2005) 57.
Grasshoff, K., Ehrhardt, M. and Kremling, K., Methods of seawater analysis, 2nd rev. and extended ed.
Verlag-Chemie, Weinheim. Germany. (1983) 419.
Horold S., Vorlop K., Tacke T. and Sell M., Development of catalysts for a selective nitrate and nitrite removal
from drinking-water, Catal. Today 17 (1993) 21.
Ilinitch O., Cuperus P., Nosova L. and Gribov E., Catalytic reduction of nitrate and nitrite ions by hydrogen:
investigation of the reaction mechanism over Pd and Pd–Cu catalysts, Catal. Today 56 (2000) 137.
Jakob M., Levanon H., Kamat P. and Nano, FTIR study of aqueous nitrat e reduction over Pd/TiO2 , Lett. 3
(2003) 353.
Jin S. and Shiraishi F., Photocatalytic activities enhanced for decompositions of organic compounds over
metalphotodepositing titanium dioxide, Chem. Eng. J. 97 (2004) 203.
Kudo A., Photocatalytic materials for water splitting, Catalysis Survey from Asia 7 (2003) 31.
Marchesini F., Irusta S., Querini C. and Miró E., Spectroscopic and catalytic characterization of Pd–In and
Pt–In supported on Al2 O3 and SiO2 , active catalysts for nitrate hydrogenation, Applied Catalysis A:
General 348 (2008) 60.
Marchesini F., Picard N. and Miró E., Study of the interactions of Pd,In with SiO2 and Al2 O3 mixed supports
as catalysts for the hydrogenation of nitrates in water, Cat. Comm. 21 (2012) 9.
Matatov-Meytal Y., Barelko V., Yuranov I. and Sheintuch M., Catalytic abatement of water pollutants, Appl.
Catal. B: Environmental 27 (2000) 127.
Abatement of nitrate in drinking water 177

Matosa J., Marino T., Molinari R. and García H., Hydrogen photoproduction under visible irradiation of
Au–TiO2 /activated carbon, Appl. Catal. A: Gen. 417–418 (2012) 263.
Nada A., Hamed H., Barakat M., Mohamed N. and Veziroglo T., Enhancement of photocatalytic hydrogen
production rate using photosensitized TiO2 /RuO2 .MV2+, International Journal of Hydrogen Energy, 33
(2008) 3264.
Ortiz de la Plata G., Alfano O. and Cassano A., Optical properties of goethite catalyst for heterogeneous
photo-fenton reactions. Comparison with titanium dioxide catalyst, Chem. Eng. J. 137 (2008) 396.
Pintar A. and Batista J., Catalytic hydrogenation of aqueous nitrate solutions in fixed-bed reactors, Catal.
Today 53 (1999) 35.
Prüsse U., Hahnlein M., Daum J. and Vorlop K., Improving the catalytic nitrate reduction, Catal. Today 55
(2000) 79.
Prusse, U., Daum, J., Bock, C. and Vorlop, K., Catalytic nitrate reduction: kinetic investigations, Stud. Surf.
Sci. Catal. 130C (2000) 2237–2242, Ref. CAS 133:256355, 2000:566770.
Ranjit K. and Viswanathan B., Photocatalytic degradation of aqueous organic pollutants using titania
supported periodic mesoporous silica, Photochem. Photobiol. A. Chem. 108 (1997) 73.
Rengaraj S. and Li X.Z., Enhanced photocatalytic reduction reaction over Bi+ 3 - TiO2 nanoparticles in presence
of formic acid as a hole scavenger, Chemosphere 66 (2007) 930.
Rossetti G., Albizzati E. and Alfano O., Decomposition of formic acid in a water solution employing the
photo-fenton reaction, Ind. Eng. Chem. Res. 41 (2002) 1436.
Sá J., Arteaga G., Daley R., Bernardi J. and Anderson J., Factors Influencing Hydride Formation in a Pd/TiO2
Catalyst, J. Phys. Chem. 110 (2006) 17090.
Sá J., Fernandez-García M. and Anderson J., Catal. Commun. 9 (2008) 1991.
Sakthivel S., Shankar M., Palanichamy M., Arabindoo B., Bahnemann D. and Murugesan V., Enhancement
of photocatalytic activity by metal deposition: characterization and photonic efficiency of Pt, Au and Pd
deposited on TiO2 catalyst, Water Res 38 (2004) 3001.
Satuf M., Brandi R., Cassano A. and Alfano O., Quantum efficiencies of 4-chlorophenol photocatalytic
degradation and mineralization in a well-mixed slurry reactor, Ind. Eng. Chem. Res. 46 (2007)43.
Satuf M., Brandi R., Cassano A. and Alfano O., Experimental method to evaluate the optical properties of
aqueous titanium dioxide suspensions, Ind. Eng. Chem. Res. 44 (2005) 6643.
Stock M. and Limbo A., New material for artificial photosynthesis, ultrasonic, ferroelectrics and frequency
control, IEEE Transactions 58 (2011) 1988.
Strukul G., Gavagnin R., Pinna F., Modaferri A., Perathoner A., Centi G., Marella M. and Tomaselli M.,
Catal. Today 55 (2000) 139.
Ward M., deKok T., Levallois P., Brender J., Gulis G., Nolan B., et al., Workgroup report: drinking-water
nitrate and health— recent findings and research needs, Environ Health Persp (2005) 113:1607–14.
Yang D., Feng W., Wu G., Li L. and Guan N., Nitrate hydrogenation on Pd–Cu/TiO2 catalyst prepared by
photo-deposition. Catal. Today 175 (2011) 356.
Zhang F., Jin R., Chen J., Shao C., Gao W., Li L. and Guan N., High photocatalytic activity and selectivity
for nitrogen in nitrate reduction on Ag/TiO2 catalyst with fine silver clusters, J. of Catalysis 232 (2005)
424.
Zhang, T., You, L. and Zhang, Y., Photocatalytic reduction of p-chloronitronenzene on illuminated nano-
titanium dioxide particles, Dyes and Pigments 68 (2006) 95.
This page intentionally left blank
CHAPTER 11

Photocatalytic inactivation of airborne microorganisms.


Performance of different TiO2 coatings

Silvia Mercedes Zacarías, María Lucila Satuf, María Celia Vaccari, &
Orlando Alfano

11.1 INTRODUCTION

The quality of indoor air has a fundamental importance for human health. Indoor air pollutants
include, among others, biological contaminants present as bioaerosols. The latter are mostly
composed of bacterial cells, fungal spores and bacterial spores (Stetzenbach et al., 2004). In
particular, bacterial endospores are characterized by their extreme latency, resistance to adverse
environmental conditions and lack of metabolism. It is the resistance to treatments, including
heat, desiccation, UV light and the action of harmful chemicals, the most important feature
of spores (Atrih and Foster, 1999). This characteristic makes them a useful reference tool for
decontamination tests. Moreover, bacterial endospores are frequently used to verify the efficiency
of sterilization devices (Josset et al., 2008).
Heterogeneous photocatalysis, an advanced oxidation technology that uses UV-A radiation
(315–400 nm) with a catalyst (mainly TiO2 ), is a potential alternative to address the problem
of biological contamination indoors. This technology has emerged as an effective treatment
method for air containing chemical and microbial contaminants due to its numerous advan-
tages. When TiO2 is photoexcited by UV-A, formation of electron-hole pairs occurs, with
the consequent generation of oxidative radical species. These radical species may cause fatal
damage to the microorganisms in contact with the catalyst surface. The process does not
involve any expensive chemical oxidant, and uses atmospheric oxygen in the presence of water
vapor.
The effectiveness of photocatalysis for the elimination of organic pollutants in air has been
proven for a wide variety of chemicals, as it is the case of alcohols, ketones, aromatics, nitrogen
and halogenated compounds (Hoffman et al., 1995; Peral et al., 1997; Blake, 2001; Zhao and
Yang, 2003; Vohra et al., 2009). However, the applicability of photocatalysis for the inactivation of
airborne microbial contaminants remains scarce. The pioneering work in the field of photocatalytic
gas phase disinfection was done by Goswami et al. (1997), reporting the total inactivation of
Serratia marcescens in a photocatalytic reactor with recirculation using TiO2 Degussa P-25.
More recently different studies on the application of photocatalysis to inactivate spore-forming
bacteria over TiO2 -coated surfaces have been published (Wolfrum et al., 2002; Lin and Li, 2003;
Pal et al., 2005; Vohra et al., 2005; Zhao et al., 2009; Chuaybamroong et al., 2010; Prasad
et al., 2011).
A relevant issue concerning photocatalytic disinfection is kinetic modeling. Kinetic expressions
governing the process are of utmost importance to identify the variables that influence the reaction
rate and to optimize the inactivation of microorganisms Additionally, modeling is essential for
the design of photocatalytic devices to solve real scale problems.
In this chapter, we present a comprehensive study on the photocatalytic inactivation of Bacillus
subtilis spores over UV-irradiated TiO2 films. In the first part of the chapter, a kinetic model
of the process is proposed (Zacarias et al., 2010). TiO2 was immobilized by the sol-gel method
179
180 S.M. Zacarías et al.

employing borosilicate glass plates as support. A reaction rate expression was derived from an
inactivation scheme which, although simple, still retains the essentials of the intervening events
and the influence of the most relevant operating variables upon the whole process. The rate of
photon absorption by the TiO2 catalyst, which represents one of the most important parameters
affecting the photocatalytic reaction rate, was predicted by using CFD software (Fluent 6.3).
The kinetic expression considers the effect of the photon absorption, the concentration of viable
spores, and the irradiation time on the spore inactivation rate.
In the second part of the chapter three different coating techniques with increasing number of
layers were assayed to improve the photocatalytic inactivation rate: TiO2 sol-gel, TiO2 Degussa
P-25, and composite TiO2 sol-gel/Degussa P-25. The amount of photocatalyst deposited on the
borosilicate glass plates and the fraction of energy effectively absorbed by the catalytic films were
determined. Diffuse transmittance and reflectance measurements of the coated and uncoated glass
plates and the subsequent application of the Net-radiation Method (Siegel and Howell, 2002) were
used to compute the fraction of energy absorbed by the TiO2 films. Finally, to objectively compare
the performance of the coatings, the photonic efficiency and the quantum efficiency of inactivation
were calculated and discussed.

11.2 KINETIC STUDY

11.2.1 Experimental set up and procedure


The inactivation of B. subtilis spores was carried out in a photocatalytic reactor comprised of a
UV-radiation emitting system, an irradiation compartment, and a support to hold the borosilicate
glass plates with the spore samples (Fig. 11.1). UV-radiation was provided by a set of seven
tubular black-light fluorescent lamps (YLX 8W/BLB) held above the irradiation compartment.
The irradiation compartment consists of a metallic rectangular box with a sliding removable
borosilicate glass plate on the top that isolates this compartment from the irradiation emitting
system. The internal walls of the box were covered with opaque black paint to minimize the
reflection of radiation. In the central zone of the irradiation compartment, where the radiation
flux was almost uniform, the photocatalytic plates with the spore samples were horizontally
held on a glass support. Also, a distilled water-filled Petri dish was included in the irradiation
compartment to secure a high relative humidity necessary to obtain a sustainable photocatalytic
activity. This variable and the temperature inside the irradiation compartment were measured with
a thermohygrometer (Mannix PTH8708 Thermo-Hygrometer TH Pen). Although the designed
photocatalytic reactor does not have an automated control of humidity and temperature, both
operating conditions remained approximately constant during the experiments: 75 ± 5% relative
humidity and 31 ± 2◦ C.
Four different irradiation levels (2.44, 0.90, 0.63 and 0.29 mW cm−2 ) were studied. These
values of the radiation flux were obtained by interposing neutral optical filters between the UV-
radiation emitting system and the irradiation compartment. The spectral transmittance of the
filters and the spectral emission of the lamps are depicted in Fig. 11.2.
Spores on bare borosilicate glass plates were irradiated to evaluate the photochemical inac-
tivation, whereas spores on TiO2 -coated borosilicate glass plates were employed to study the
photocatalytic inactivation. In all cases the size of the plates was 2 × 2 cm. A sol-gel technique
that uses titanium tetraisopropoxide as precursor (Yamazaki–Nishida et al., 1993) was employed
to obtain the TiO2 thin films. First, the hydrolysis of the titanium tetra-isopropoxide was con-
ducted in an acid aqueous medium. A volume of 415 mL of triple-distilled water was mixed
with 4.5 mL of concentrated nitric acid (Anedra, 65%), and 35 mL of titanium isopropoxide
(Aldrich, 97%) was added to this mixture. Under such conditions, the hydrolysis of the precursor
proceeded vigorously, producing large lumps of hydrated TiO2 . The dispersion of the lumped
particles was achieved by stirring the suspension over a period of 10 h at 80◦ C, until a clear sol
of TiO2 nano-particles was obtained. To prepare the surface of the borosilicate glass before the
Photocatalytic inactivation of airborne microorganisms. 181

Figure 11.1. Experimental setup: (a) Schematic diagram, (b) Picture of the top view. Reprinted with per-
mission from Ind. Eng. Chem. Res., Efficiency evaluation of different TiO2 coatings on the
photocatalytic inactivation of airborne bacterial spores, S. M. Zacarias, M. L. Satuf, M. C.
Vaccari, O. M. Alfano. 51, 13599–13608. Copyright 2012 American Chemical Society.

coatings, the plates were first washed with isopropyl alcohol and triple-distilled water for 20 min
under sonication, and then heated during 8 h at 500◦ C to remove any trace of organic materials.
The dip-coating technique was used to obtain the TiO2 films: the glass plates were immerse in
the clear sol of TiO2 obtained, and then removed at a withdrawal speed of 3 cm min−1 at room
temperature (25◦ C). Subsequently, the plates were dried in an oven at 80◦ C and then heated at
200C for 6 h to increase the adherence of the thin film and to induce the formation of the anatase
phase, which is the most active crystalline form of TiO2 for photocatalytic applications.
182 S.M. Zacarías et al.

Figure 11.2. Spectral emission of the lamps (—) and spectral transmittances of the neutral filters: 37%
T (· · · ), 26% T (- - -) and 12.5% T (– –). Adapted from Journal of Photochemistry and
Photobiology A: Chemistry, 214, S. M. Zacarias, M. C. Vaccari, O. M. Alfano, H. A. Irazoqui,
G. E. Imoberdorf, Effect of the radiation flux on the photocatalytic inactivation of spores of
Bacillus subtilis, 171–180, Copyright 2010, with permission from Elsevier.

Bacillus subtilis spores (ATCC 6633 strain) were adopted as the model microorganism. Fol-
lowing the technique proposed by Shehata and Collins (1972), suspensions of spores in distilled
water were obtained. A sporulation medium (nutritive agar from Britania Laboratories, with
0.05% MnSO4 and 0.05% MgSO4 ) contained in a Roux bottle was inoculated and incubated at
30◦ C during a 7- to 10-day period. Next, the surface of the sporulation medium was washed
with sterile 0.9% NaCl solution to recover the spores and vegetative cells. Then, the cells’ sus-
pension was centrifuged three times during 15 minutes at 3500 rpm. To induce vegetative cell
lysis, the suspension was kept at 30◦ C during 48 h. After this period, according to the instructions
outlined above, the vegetative cells and the spores were washed again, repeating this procedure
three times. The spores’ suspension was finally suspended in sterile distilled water and conserved
at 4◦ C. The concentration of spores in the suspension was comprised between 1.0 × 106 and
1.0 × 107 CFU cm−3 .
For the irradiation experiments, to eliminate any remaining vegetative cell, aliquots of 1 mL of
the suspension of spores were kept at 80◦ C during 10 min. Then, the glass plates coated with the
TiO2 films and the bare glass plates were covered with aliquots of 10 µL of the spores’ suspen-
sion. Next, the samples were dried for 1 hour at 30◦ C, and then introduced into the irradiation
compartment. Subsequently, the dried plates containing the spores were exposed to a given irra-
diation level for different time periods (6, 12, 24 or 48 h). Finally, the samples were removed
from the irradiation compartment and the remaining viable spores were counted. For this pur-
pose, each plate was placed in a tube with 10 mL of sterile extraction solution (0.1% peptone in
sterile distilled water), and then gently scraped using a small sterile spatula. The solution was
shaken at 200 rpm for 15 min. Aliquots of 1 mL were spread onto agar plates, incubated for 48 h
at 30◦ C, and the colony forming units (CFU) were counted. When necessary, these aliquots were
conveniently diluted with the sterile extraction solution before the quantification step to obtain 30
to 300 CFU per agar plate. The detection limit of the quantification procedure is 6.7 × 102 CFU
Photocatalytic inactivation of airborne microorganisms. 183

Figure 11.3. Inactivation curves for different irradiation times of spores irradiated on bare plates () and
spores irradiated on TiO2 coated plates for an incident radiation flux of 2.44 mW cm−2 (♦),
and 0.63 mW cm−2 ().

cm−2 The tests were repeated twice for each experimental condition studied, and each repetition
was made in duplicate.

11.2.2 Inactivation of spores


The experimental results obtained when B. subtilis spores were exposed to UV-A radiation are
shown in Figure 11.3. The figure presents results obtained with bare plates and with TiO2 coated
plates under two irradiation levels: 2.44 and 0.63 mW cm−2 .
N0 represents the initial concentration of spores, and N expresses the remained viable spores
after irradiation. Trendlines were plotted to facilitate observation. When the spores were irradiated
on the surface of glass plates without TiO2 , inactivation was not significant. On the contrary, the
viability of spores notably diminished with the irradiation time when they were exposed to UV-A
radiation over theTiO2 film. Also, the inactivation extent increased with the radiation flux reaching
the photocatalytic plate.
It is worth noting that no detectable changes in the survival of spores were observed when TiO2
films were kept in the dark.

11.2.3 Kinetic modeling


11.2.3.1 Proposed kinetic model
The exact mechanism of the photocatalytic inactivation of microorganisms is still under debate.
Sun et al. (2003) and Cho et al. (2005) worked with Escherichia coli in water disinfection, and
they proposed the holes (h+ ) and HO• radicals as those responsible for the inactivation of this
microorganism. In this study, a simplified scheme to represent the photocatalytic inactivation
of B. subtilis spores is proposed in Table 11.1. The first reaction represents the semiconductor
activation by the absorption of UV photons, and the generation of holes (h+ ) and electrons (e− ).
The photo-generated h+ may interact with water molecules to produce HO• radicals (reaction 2),
while free e− may be transferred to adsorbed molecular oxygen thus generating • O− 2 radicals
(reaction 3). Subsequently, the superoxide radical anion reacts to form hydrogen peroxide or
peroxide anion (Turchi and Ollis, 1990; Legrini et al., 1993). Also, an HO• radical may react
184 S.M. Zacarías et al.

Table 11.1. Simplified inactivation scheme. Reprinted from Journal of Photochem-


istry and Photobiology A: Chemistry, 214, S. M. Zacarias, M. C. Vaccari,
O. M. Alfano, H. A. Irazoqui, G. E. Imoberdorf, Effect of the radia-
tion flux on the photocatalytic inactivation of spores of Bacillus subtilis,
171–180, Copyright 2010, with permission from Elsevier.

N◦ Reaction Kinetic expressions

(1) TiO2 + hν → TiO2 + h+ + e− r1 =


ea,s
(2) h+ + H2 O(ads) → HO• + H+ r2 = k2 [h+ ] [H2 O](ads)
(3) e − + O 2 → • O−
2 r3 = k3 [e− ] [O2 ]
(4) 2 HO• → H2 O2 r4 = 2k4 [HO• ]2
(5) +
h + Surface Compounds → Products r5 = k5 [h+ ] [SC]
(6) HO• + Surface Compounds → Products r6 = k6 [HO• ] [SC]
(7) HO• + Viable Spores → Inactive Spores r7 = k7 [HO• ] [N]
(8) h+ + e− → Thermal Energy r8 = k8 [h+ ] [e− ]

with other HO• radical to form H2 O2 (reaction 4). Nevertheless, such event is very unlikely to
occur. So, we may assume that reaction 4 will progress at a very slow rate and the concentration
of H2 O2 will be negligible. Within the assumptions made for modeling the inactivation process,
we assumed that the spore is composed of a single coat, including the inner and outer coat, and
the cortex of the spore. Reaction 5 represents the interaction between photo-generated h+ and the
compounds of the spore coat. Reaction 6 represents the progressive photocatalytic degradation
of the compounds of the spore coat at its point of contact with the photocatalytic surface, which
does not cause lethal injuries to the spores. Step 7 represents the loss of spore viability to obtain
inactive spores. Inactive spores are assumed to differ from the active ones only in their ability
to germinate. The last reaction (8) represents the recombination of the photogenerated holes and
electrons with the resulting loss of thermal energy.
It is assumed that, before the irradiation process begins, all the spores are intact, and when
irradiation starts, the photogenerated HO• radicals and h+ begin to progressively oxidize the
organic compounds of the spore’s coat. Moreover, we propose that the damage in the coat and in
the core is accumulative. Therefore, the initially intact spores evolve through states of increasing
degradation extent. At some point, this damage reaches a level that cannot be repaired or reversed
by the spores, and then the spore is inactive.
According to step (7), the rate of inactivation of the spores is assumed proportional to the spore
concentration and to the [HO• ] radical species concentration:
d[N]
= −k7 [HO• ][N] (11.1)
dt
By applying the micro steady state approximation to holes (h+ ), electrons (e− ), and HO•
radicals we get (Zacarías et al., 2010):

+ k3 [O2 ]ads 4 k8
ea,s
[h ] = 1+
−1 (11.2)
2 k8 k3 [O2 ]ads k2 [H2 O]ads + k5 [SC]




k2 [H2 O]ads + k5 [SC] 4 k8
ea,s
[e ] = 1+
−1 (11.3)
2 k8 k3 [O2 ]ads k2 [H2 O]ads + k5 [SC]

k [H O] k [O ] 4 k
e a,s
[HO• ] =
2 2 3 2 8
ads
1+
−1 (11.4)
2k8 (k6 [SC] + k7 [N]) k3 [O2 ]ads k2 [H2 O]ads + k5 [SC]
Photocatalytic inactivation of airborne microorganisms. 185

The chemical compounds of the spores’coat in direct contact with theTiO2 films are represented
as surface compounds (i.e., [SC]). These compounds will act as scavengers of HO• , since they
will reduce the availability of HO• to react with the vital targets of the spores, protecting them
from being inactivated. Under the operating conditions utilized in our experiments, it is very
unlikely that complete mineralization of these surface compounds would take place. Hence, no
substantial variations of its concentration are expected and the concentration of surface compounds
was considered constant. Additionally, the concentration of adsorbed water and oxygen on the
catalytic surface is assumed to remain constant during the irradiation process.
Then, introducing the resulting expression of [HO• ] into Equation (1) and integrating, we arrive
to the following equation
   
[N] = [N0 ] exp K3 ([N0 ] − [N]) − K1 K2 + ea,s − K2 t (11.5)


k2 k3 [O2 ]ads [H2 O]ads 1 k3 [O2 ]ads k2 [H2 O]ads + k5 [SC]
where K1 = k7 K2 , K 2 =
2k8 k6 [SC] 4 k8

k7
and K3 =
k6 [SC]
The final set of adjustable kinetic parameters is {K1, K2, K3}.

11.2.3.2 Radiation model


Photon absorption by the TiO2 film is the initial step of the photocatalytic inactivation. The
kinetics of this process depends on the Local Superficial Rate of Photon Absorption (LSRPA
or ea,s ) (Imoberdorf et al., 2005). Because the ea,s values cannot be experimentally measured,
a radiation field model was devised, which takes into account the radiation emission from the
lamps, its propagation in the irradiation compartment, its attenuation in the borosilicate glass at
the top of this compartment, and finally the absorption of photons in the TiO2 films. A more
detailed description of the radiation model can be found in Zacarías et al. (2010).
The radiative transfer equation (RTE) was solved by means of the Discrete Ordinate Method
(DOM) using the Fluent Software (Fluent 6.3). Figure 11.4 shows, as a gradation of grays, the
values of the computed radiation flux incident at the different nodes of the plane where the spore
samples were located; note that the darker zones correspond to the less illuminated regions.
A radiometer (IL 1700 – SED005 – WBS 320) was used to get local measurements of the
radiation flux incident at different positions on a plane. These experimental values were compared
with the predicted values of the local radiation flux obtained with CFD simulations. The agreement
between simulated and experimental results was satisfactory, with an average relative error of 11%.
From Fig. 11.4, significant variations can be observed at different locations of the irradiation
compartment. To ensure that all samples are exposed to the same irradiation flux, only the most
illuminated surface at the central zone of the irradiation compartment was used. In this zone,
an acceptable difference between the most illuminated site on the sample plate and the less
illuminated one was obtained (<17%).
By setting up local radiative energy balances in terms of local energy fluxes, the LSRPA was
computed at each point of the catalytic film (Imoberdorf et al., 2005):
ea,s (x) = qf ,in (x) − qf ,tr (x) (11.6)

In Equation (6), qf ,in (x) stands for the local radiative flux that reaches the catalytic surface and
qf ,tr (x) the local radiative energy flux that is transmitted through the TiO2 films. By subtracting
the radiation attenuation in the TiO2 films from the radiation flux reaching those films, the rate
of photon absorption can be calculated as follows:
  
ea,s (x) = qf ,in (x) 1 − exp −κλ,f λlamp ef (11.7)

where κλ,f is the absorption coefficient of the TiO2 film and ef represents the film thickness.
186 S.M. Zacarías et al.

Figure 11.4. Radiative energy flux on the irradiation compartment cross section represented by a gradation
of grays. The heavy-bordered rectangle at the center represents the zone employed to locate
the sample plates. Reprinted from Journal of Photochemistry and Photobiology A: Chemistry,
214, S. M. Zacarias, M. C. Vaccari, O. M. Alfano, H. A. Irazoqui, G. E. Imoberdorf, Effect
of the radiation flux on the photocatalytic inactivation of spores of Bacillus subtilis, 171–180,
Copyright 2010, with permission from Elsevier.

11.2.3.3 Kinetic parameters estimation


The complete set of experimental results was employed to fit the kinetic parameters of Equation
(11.5). The objective function (OF) to be minimized was defined as the summation over all
experimental points of the squared differences between each experimental value of N and the
corresponding value calculated with the kinetic expressions divided by the standard deviation of
the experimental value:

n
([N ]exp,i − [N ]mod,i )2
OF(K1 , K2 , K3 ) = (11.8)
i=1
σi2

The set (K1 , K2 , K3 ) of kinetic parameters values that minimizes the OF was considered the
best-fit. In order to find this value, the genetic algorithm (GA) was used. In general terms, the GA
is an adaptive heuristic search algorithm inspired on evolution, natural selection, and genetics.
Photocatalytic inactivation of airborne microorganisms. 187

Figure 11.5. Comparison between model predictions and experimental values of viable spores obtained
after the irradiation process. Reprinted from Journal of Photochemistry and Photobiology A:
Chemistry, 214, S. M. Zacarias, M. C. Vaccari, O. M. Alfano, H. A. Irazoqui, G. E. Imoberdorf,
Effect of the radiation flux on the photocatalytic inactivation of spores of Bacillus subtilis,
171–180, Copyright 2010, with permission from Elsevier.

This minimization method was selected because it is capable of finding the absolute minimum of a
given OF rather than its local minima, as most of the alternative methods do. The parameter values
obtained with the GA for equation (5) are: K1 = 123.76 cm Einstein−1/2 h−1/2 , K2 = 2.53 × 10−9
Einstein h−1 cm−2 , and K3 = 1.95 × 10−27 cm2 CFU−1 .
Analyzing the values of the kinetic parameters, it is possible to conclude that the terms affected
by K2 and K3 can be neglected in Equation (5). Then, the final kinetic equation is:

[N ] = [N0 ]exp(−K1 ea,s t) (11.9)

From the above equation, it can be seen that the inactivation rate has a square root dependence
on the absorbed radiation by the photocatalytic film (ea,s ).
Figure 11.5 compares the experimental results of the inactivation of the spores with those
predicted by the simplified kinetic model (Equation 11.9) for all the experimental data obtained
in this work. Good agreement is observed.

11.3 STUDY OF DIFFERENT TiO2 COATINGS

The photocatalytic experiments with three different TiO2 coatings were conducted in the photocat-
alytic reactor detailed in Section 11.2.1., with two minimum modifications: (1) a saturated solution
of ammonium sulfate was included in the irradiation compartment to control more precisely the
relative humidity percentage; (2) the closure system of the reactor was improved. Throughout this
second set of experiments, the relative humidity and temperature were kept constant at 70% and
40◦ C, respectively (Zacarías et al., 2012).
Employing a similar procedure to the one described in the first section, the plates with Bacil-
lus subtilis spores were placed in the irradiation compartment and exposed to UV-A radiation.
Nevertheless, the irradiation times with these different coating techniques and with increasing
number of layers were considerably shorter: 2, 4, 6 and 8 h.
188 S.M. Zacarías et al.

The coatings were characterized by SEM photographs. In addition, the amount of TiO2 immo-
bilized on the plates and the fraction of energy absorbed by the films were also determined.
Finally, to objectively compare the performances of the coatings, the photonic efficiency ηapp ,
also known as apparent quantum efficiency, and the quantum efficiency of inactivation ηabs were
evaluated.
A spectrophotometric procedure adapted from Jackson et al.(1991) was used to quantify the
amount of TiO2 deposited on the glass plates. This method involves the digestion of the catalyst
followed by a colorimetric detection.
The fraction of incident radiation effectively absorbed by the films can be expressed as
αf ,λ φλ , where αf ,λ is the fraction of energy absorbed by the TiO2 film at wavelength λ, and
φλ is a normalized distribution function of the wavelengths that reach the coated plates. Here φλ
takes into account the spectral emission of the lamps and the slight modification of the lamps
spectrum produced by the absorption of the borosilicate glass that separates the lamps from the
irradiation compartment, mainly at λ lower than 340 nm.
To compute αf ,λ , the Net-radiation Method was employed (Siegel and Howell, 2002). This
method considers energy absorption, reflection and transmission produced by multiple parallel
layers. The fractions of incident energy transmitted (T ), reflected (R) and absorbed (α) are given
by the expressions:
Tf ,λ Tg,λ
Tfg,λ = (11.10)
1 − Rf ,λ Rg,λ

Rg,λ Tf2,λ
Rfg,λ = Rf ,λ + (11.11)
1 − Rf ,λ Rg,λ

αf ,λ = 1 − Tf ,λ − Rf ,λ (11.12)

where the subscripts f, g and fg represent film, glass and film+glass optical properties, respec-
tively. By spectrophotometric measurements of diffuse transmittance and reflectance of the bare
and coated plates, the spectral values of Tg , Rg , Tfg and Rfg were obtained. Then, with Equations
(11.10) and (11.11), we could obtain Tf and Rf , and, finally, calculate αf (Equation 11.12) within
the emission range of the lamps (300 to 400 nm). A detailed description of the measurement
procedure can be found elsewhere (Zacarías et al., 2012).

11.3.1 Efficiency parameters


ηapp can be expressed as the ratio of the initial photocatalytic inactivation rate to the rate of radiation
energy reaching the plate covered with the TiO2 film, over the entire range of wavelengths under
consideration:
  
− dN dt |t=0
ηapp =   (11.13)
qf ,in
Airr

The initial inactivation rate (dN /dt)|t=0 was calculated from experimental data. To do this, the
decay of viable spores as a function of the irradiation time was fitted with the exponential equation
[N ] = [N0 ]e(−kt) where k represents the apparent kinetic constant. Then, the first derivative of the
exponential equation at time zero was obtained.
The denominator of Equation (11.13) represents the incident radiation flux averaged over the
irradiated area Airr . It was experimentally measured with a radiometer and the following value was
obtained: 1.89 mW cm−2 , which is equivalent to 2.01 × 10−5 Einstein cm−2 h−1 (or 1.21 × 1019
photon cm−2 h−1 ).
Photocatalytic inactivation of airborne microorganisms. 189

ηabs can be defined as the ratio of the initial photocatalytic inactivation rate to the rate of the
absorbed radiation energy by the TiO2 film, within the useful wavelength range:
  

− dN dt t=0
ηabs =   (11.14)
efa,s
Airr

where efa,s Airr


is the local superficial rate of photon absorption by the catalytic film, averaged
over the irradiated area. In turn, efa,s Airr can be expressed as (Briggiler Marcó et al., 2011):

efa,s A = qf ,in A − qf ,tr A − qf ,rf A = qf ,in A αf ,λ ϕλ (11.15)
irr irr irr irr irr
λ

where qf ,in is the local radiative flux that reaches the coated plate, qf ,tr the local radiative flux
transmitted through the catalytic plate, and qf ,rf the local radiative flux reflected by the TiO2
surface.
Introducing Equation (11.15) into (11.14) yields
  

− dN dt t=0
ηabs =  (11.16)
qf ,in A αf ,λ ϕλ
irr
λ

11.3.2 Preparation of the photocatalytic coatings


The first coating procedure (TiO2 sol-gel) consists in the sol-gel method detailed in Section
11.2.1. The second one (TiO2 Degussa P-25) employs a suspension of 150 g L−1 of Degussa
P-25 TiO2 powder (Evonik Degussa) in ultrapure water at a pH value of 1.5, adjusted with
HNO3 (van Grieken et al., 2009). The third coating (sol-gel/Degussa P-25) was obtained by a
composite sol-gel/Degussa P-25 TiO2 technique (Keshmiri et al., 2004). Denatured ethanol (85%
ethanol + 15% methanol) was mixed with ultrapure water and concentrated HCl. After a few
minutes of agitation, titanium tetraisopropoxide (Sigma Aldrich, 97%) was added to the mixture.
Hydrolysis was carried out by drop wise addition of the precursor to the prepared solution, while
stirring. This mixture was agitated for 2 h. After this period, TiO2 Degussa P-25 was added and
stirred for 12 h.
To prepare the borosilicate glass for the coating procedure, the plates were washed with a
solution containing 20 g of potassium hydroxide, 250 mL of isopropyl alcohol and 250 mL of
ultrapure water. The plates remained in contact with the washing solution for 24 h, and then
2 h under sonication (Ultrasonik 300). Next, they were heated during 8 h at 500◦ C. In the three
coatings assayed, the TiO2 immobilization over the glass plates was made by the dip-coating
technique, with a withdrawal speed of 3 cm min−1 at room temperature (25◦ C). Nevertheless,
after the corresponding coating was made, each catalyst was treated differently. For the TiO2
sol-gel coating, the plates were dried in an oven at 80◦ C for 1 h and then heated at 200◦ C for 6 h.
For the coatings with TiO2 Degussa P-25, the plates were dried in an oven at 110◦ C for 24 h and
then heated at 500◦ C for 2 h, with a heating rate of 5 ◦ C min−1 . Finally, for the sol-gel/Degussa
P-25 coating, the plates were dried at room temperature (25◦ C) for 1 h and then heated at 500 ◦ C
for 2 h, with a heating rate of 11◦ C min−1 .
To obtain two or three layers of catalyst on the glass plates, after the drying and heating steps, the
next coating was made by repeating the procedure followed to deposit the first layer. In particular,
it was not possible to obtain a third layer with the sol-gel/Degussa P-25 technique due to the poor
adhesive properties of the retrieved film.

11.3.3 Characterization of the photocatalytic plates


Table 11.2 reports the amount of TiO2 deposited on the plates and the absorbed fraction by the
films. For the three coating techniques, a gradual increase in the TiO2 content [mg cm−2 ] is
190 S.M. Zacarías et al.

Table 11.2. Characterization of the coatings.

sol-gel Degussa P-25 sol-gel/Degussa


Layers Layers Layers
1 2 3 1 2 3 1 2

TiO −2
 2 [mg cm ] 0.05 0.11 0.16 1.26 1.77 2.13 0.58 2.01
αf ,λ ϕλ 0.281 0.284 0.337 0.879 0.894 0.891 0.829 0.915

Figure 11.6. SEM photographs of the TiO2 coating surfaces with two layers. (a) sol-gel; (b) Degussa
P-25; (c) and (d) sol-gel/Degussa. The white bar represents: 1 µm in (a)-(c), and 10 µm in (d).
Reprinted with permission from Ind. Eng. Chem. Res., Efficiency evaluation of different TiO2
coatings on the photocatalytic inactivation of airborne bacterial spores, S. M. Zacarías, M. L.
Satuf, M. C. Vaccari, O. M. Alfano. 51, 13599–13608. Copyright 2012 American Chemical
Society.

observed with the number of layers. The sol-gel films contain the lowest amounts of catalyst,
with the resulting lowest values of absorbed radiation. On the contrary, the coatings made with
Degussa P-25 and sol-gel/Degussa present considerably higher amounts of TiO2 and high fractions
of absorbed radiation.
SEM images of the different coatings with two layers are shown in Fig. 11.6. The coating
obtained with the sol-gel method presents the thinnest and most uniform surface. The Degussa
P25 image reveals a more irregular surface, where dark regions of uncoated glass are observed.
The roughest surface was obtained with the composite technique, with considerable irregularities
throughout the plate. These inhomogeneities are better observed in Fig. 11.6(d), with less magni-
fication. It should be noted that the adherence of the composite coating to the glass was also poor.

11.3.4 Evaluation of photocatalytic efficiencies


The experimental results of the photocatalytic inactivation of Bacillus subtilis spores over the
coated plates and the curve fittings are shown in Fig. 11.7. It is important to mention that no
detectable changes in the survival of spores were observed when the samples were kept in the
dark or in absence of catalyst.
Photocatalytic inactivation of airborne microorganisms. 191

Figure 11.7. Spore inactivation curves for the three coating techniques. Experimental values: ♦, Degussa
P-25; , sol-gel/Degussa P-25; , sol-gel. Fitting (—, - - - , · · · ). Reprinted with permission
from Ind. Eng. Chem. Res., Efficiency evaluation of different TiO2 coatings on the photocat-
alytic inactivation of airborne bacterial spores, S. M. Zacarías, M. L. Satuf, M. C. Vaccari,
O. M. Alfano. 51, 13599–13608. Copyright 2012 American Chemical Society.

The initial inactivation rates (numerator of Equations (11.13) and (11.14)), calculated from the
slope of the inactivation curves, are shown in Table 11.3. This table also reports the values of
the energy absorbed by the coated plates (denominator of Equation (11.14)). The denominator of
Equation (11.13) (incident radiation flux) was always 1.21 × 1019 photon cm−2 h−1 .
Figure 11.8 shows the results of both photocatalytic efficiency parameters, for each catalyst
studied and for the different layers assayed.
192 S.M. Zacarías et al.

Table 11.3. Initial inactivation rates and energy absorbed by each coating. Adapted with permission from
Ind. Eng. Chem. Res., Efficiency evaluation of different TiO2 coatings on the photocatalytic
inactivation of airborne bacterial spores, S.M. Zacarías, M.L. Satuf, M.C. Vaccari, O.M. Alfano.
51, 13599–13608. Copyright 2012 American Chemical Society.

–dN/dt|t=0 [CFU cm−2 h−1 ] × 10−5 efa,s  [photon cm−2 h−1 ] × 10−19

Layers Layers
Coating 1 2 3 1 2 3

sol-gel 0.40 0.47 0.49 0.34 0.34 0.40


Degussa P-25 1.40 2.45 1.42 1.07 1.08 1.08
sol-gel/Degussa 0.23 0.69 — 1.00 1.11 —

Figure 11.8. Photocatalytic inactivation efficiencies.

11.3.5 Discussion
Concerning the coatings performed with the sol-gel technique, an expected gradual increase in
the TiO2 content with the number of layers was found, resulting in an increment in the absorption
of radiation, inactivation rate and photonic efficiency. However, regarding ηabs , the sample with
two layers was the most efficient.
Photocatalytic inactivation of airborne microorganisms. 193

With respect to the Degussa coating, a high percentage of the incident radiation was absorbed
by the three samples tested, as evidenced in Table 11.2. Besides, an appreciable increment in the
amount of immobilized TiO2 was found as the number of layers increased. It is noted that the
highest inactivation rate was obtained with two layers. Because the amount of absorbed energy
is similar for the three samples, the quantum efficiency is also higher for the plate with two
layers of TiO2 . In fact, the highest quantum efficiency of all tested samples was obtained with the
Degussa coating with two layers: ηabs = 2.27 × 10−14 (Fig. 11.8(b)). With three layers, although
the amount of catalyst is higher than the quantity obtained with two layers, the inactivation rate
is lower and similar to the one found with one layer. A possible explanation can be found in
the surface roughness of the retrieved film with three layers. Bacterial spores could lay in the
irregularities of the coating where radiation could not arrive (Caballero et al., 2009).
It is worth to highlight that, although the amount of catalyst and the percentage of radiation
absorption with the Degussa coating are significantly higher than the ones obtained with the
Sol-gel coating, the values of the quantum efficiency are somewhat similar.
Finally, a significant increment in the amount of immobilized TiO2 was found with the sol-
gel/Degussa technique between one and two layers. An increase in the radiation absorption,
inactivation rate and efficiency parameters with the number of layers were also obtained. Even
though the coatings obtained with the sol-gel/Degussa technique contain high amounts of TiO2 and
absorb a high percentage of the incident radiation, the inactivation rates were similar to the ones
obtained with the sol-gel samples, and the values of ηabs are the lowest as compared with the other
two coating techniques. The explanation to this behavior can be ascribed to the irregularities of the
film, as exposed in the case of three layers with Degussa. Also, a deficiency in the contact between
the spores and the film could explain the low inactivation efficiency of the composite coatings.

11.4 CONCLUSIONS

Significant reduction of the viability of Bacillus subtilis spores over UV-irradiated TiO2 films has
been achieved. Besides, the proposed methodology for kinetic modeling allows the simulation of
the photocatalytic inactivation process. It was found that the kinetic model was able to represent
satisfactorily the inactivation of spores under different irradiation conditions. Although the results
are encouraging, further research is needed to construct useful photocatalytic devices for air
disinfection.
The calculation of efficiency parameters represents a valuable tool to compare different coating
techniques with increasing number of TiO2 layers, to select the best alternative and to improve
the design of disinfection systems. To achieve this goal, the photonic efficiency ηapp and the
quantum efficiency of inactivation ηabs can be computed. Nevertheless, ηabs provides more useful
information because it shows how efficiently the absorbed radiation is employed to inactivate
microorganisms, giving an indirect insight into the photocatalytic mechanism. Also, it can help
to design better coatings, as the results can evidence deficient contact between bacteria and
photocatalyst, or poor absorption of radiation.

ACKNOWLEDGEMENTS

Section 11.2 of this chapter corresponds to a contribution that has been written with the following
co-workers: H.A. Irazoqui and G.E. Imoberdorf.
The authors would like to thank Universidad Nacional del Litoral (UNL), Consejo Nacional de
Investigaciones Científicas y Técnicas (CONICET) and Agencia Nacional de Promoción Cientí-
fica y Tecnológica (ANPCyT) for their financial support. Thanks are also given to Eng. Gerardo
Rintoul and Antonio C. Negro, for their participation in some parts of the experimental work, and
to Professor María C. Lurá for her valuable support.
194 S.M. Zacarías et al.

NOMENCLATURE

ATCC American Type Culture Collection


CFU Colony Forming Units
e− electrons
ef thickness of the TiO2 films (cm)
efa,s superficial rate of photon absorption by the TiO2 films (photon cm−2 h−1 )
h+ positives “holes”
k kinetic constant (h−1 )
ki kinetic parameters, units depends on the reaction step
Ki global kinetic parameters, units depends on the reaction step
N viable spore concentration (CFU cm−2 )
N0 initial viable spore concentration (CFU cm−2 )
HO• hydroxyl radical
• −
O2 superoxide radical
qf ,in local radiative flux that reaches the TiO2 films (photon cm−2 h−1 )
qf ,tr local radiative flux transmitted through the TiO2 films (photon cm−2 h−1 )
qf ,rf local radiative flux reflected by the TiO2 films (photon cm−2 h−1 )
ri reaction rate (mol s−1 cm−2 )
R reflectance (dimensionless)
SCi spore’s surface compounds
t time (h)
T transmittance (dimensionless)
Greek letters

primary quantum yield


κλ,g λlamp absorption coefficient of the borosilicate glass (cm−1 )
κλ,f λlamp absorption coefficient of the TiO2 film (cm−1 )
αf fraction of energy absorbed by the TiO2 films (dimensionless)
η photocatalytic inactivation efficiency (CFU photon1 )
ϕλ distribution function of the wavelengths that reach the TiO2 catalytic
films (dimensionless)
Special symbols
[] concentration
 average
Subscripts
λ relative to a specific wavelength
abs denotes the quantum efficiency of inactivation
app denotes the photonic efficiency (or apparent quantum efficiency)
Airr relative to the irradiated area
f relative to a property of the TiO2 film
fg relative to a property of the TiO2 film + glass
g relative to a property of the bare borosilicate glass plate

REFERENCES

Atrih, A. & Foster, S. J: The role of peptidoglycan structure and structural dynamics during endospore
dormancy and germination. Antonie van Leeuwenhoek, 75 (1999), pp. 299–307.
Blake, D.M.: Bibliography of work on the heterogeneous photocatalytic removal of hazardous compounds
from water and air 2001. National Renewable Energy Laboratory, Golden, Colorado. 2001.
Briggiler Marcó, M., Quiberoni, A. L., Negro, A. C., Reinheimer, J. A. & Alfano, O. M.: Evaluation of the
photocatalytic inactivation efficiency of dairy bacteriophages. Chem. Eng. J., 172 (2011), pp. 987–993.
Photocatalytic inactivation of airborne microorganisms. 195

Caballero, L., Whitehead, K. A., Allen, N. S. & Verran, J.: Inactivation of Escherichia coli on immobilized
TiO2 using fluorescent light. J. Photochem. Photobiol., A., 202 (2009), pp. 92–98.
Cho, M. Chung, H. Choi & W. Yoon, J.: Different inactivation behaviors of MS-2 phage and Escherichia coli
in TiO2 photocatalytic disinfection. Applied and Environmental Microbiology, 71 (2005), pp. 270–275.
Chuaybamroong, P., Chotigawin, R., Supothina, S., Sribenjalux, P., Larpkiattaworn, S. & Wu, C.-Y.: Efficacy
of photocatalytic HEPA filter on microorganism removal. Indoor Air, 20 (2010), pp. 246–254.
Goswami, D. Y., Trivedi, D. M. & Block, S. S.: Photocatalytic disinfection of indoor air. J. Sol. Energ. Eng
119 (1997), pp. 92–96
Hoffman, M.R., Martin, S.T., Choi, W. & Bahnemann, D.W.: Environmental applications of semiconductor
photocatalysis. Chemical Review, 95 (1995), pp. 69–96.
Imoberdorf, G. Irazoqui, H. A. Cassano, A. E. & Alfano, O. M.: Photocatalytic degradation of tetra-
chloroethylene in gas phase on TiO2 films: A kinetic study Ind. Eng. Chem. Res., 44 (2005), pp.
6075–6085.
Jackson, N. B., Wang, C. M., Luo, Z., Schwitzgebel, J., Ekerdt, J. G., Brock, J. R. & Heller, A.: Attachment
of TiO2 Powders to Hollow Glass Microbeads: Activity of the TiO2 -Coated Beads in the Photoassisted
Oxidation of Ethanol to Acetaldehyde. J. Electrochem. Soc., 138 (1991), pp. 3660–3664.
Josset, S. Keller, N. Lett, M.-C. Ledoux M. J. & Keller, V.: Numeration methods for targeting photoactive
materials in the UV-A photocatalytic removal of microorganisms. Chem. Soc. Rev., 37 (2008), pp. 744–755
Keshmiri, M., Mohseni, M. & Troczynski, T.: Development of novel TiO2 sol–gel-derived composite and its
photocatalytic activities for trichloroethylene oxidation. Appl. Catal., B., 53 (2004), pp. 209–219.
Legrini, O., Oliveros, E. & Braun, A. M.: Photochemical processes for water treatment. Chem. Rev. 93
(1993), pp. 671–698.
Lin, C. Y. & Li, C. S. Inactivation of microorganisms on the photocatalytic surfaces in air. Aerosol Sci.
Technol 37 (2003), pp. 939–946
Pal, A., Min, X., Yu, L. E., Pehkonen, S. O. & Ray, M.: Photocatalytic inactivation of bioaerosols by TiO2
coated membrane. Int. J. Chem. React. Eng. 3:A45 (2005), pp. 1–12.
Prasad, G. K., Ramacharyulu, P. V. R. K., Merwyn, S., Agarwal, G. S., Srivastava, A. R., Beer Singh,
Rai, G. P. & Vijayaraghavan, R.: Photocatalytic inactivation of spores of Bacillus anthracis using titania
nanomaterials. J. Hazard. Mater., 185 (2011), pp. 977–982.
Peral, J., Domènech, X. & Ollis, D.F.: Heterogeneous photocatalysis for purification, decontamination and
deodorization of air. Journal of Technology and Biotechnology, 70 (1997), pp. 117–140.
Shehata, T. E. & Collins, E. B.: Sporulation and heat resistance of psychrophilic strains of Bacillus. Journal
of Dairy Science, 55 (1972), pp. 1405–1409.
Siegel, R. & Howell, J.: Thermal Radiation Heat Transfer, 4th ed., Taylor and Francis, New York, 2002.
Stetzenbach, L. D., Buttner, M. P. & Cruz, P.: Detection and enumeration of airborne biocontaminants.
Current Opinion in Biotechnology, 15 (2004), pp. 170–174.
Sun, D. D., Tay, J. H. & Tan, K. M. Photocatalytic degradation of E. coliform in water. Water Research, 37
(2003), pp. 3452–3462.
Turchi, C. S. & Ollis, D. F.: Photocatalytic degradation of organic water contaminants: mechanisms involving
hydroxyl radical attack. J. Catal. 122 (1990), pp. 178–192.
van Grieken, R., Marugán, J., Sordo, C. & Pablos, C.: Comparison of the photocatalytic disinfection of E.
coli suspensions in slurry, wall and fixed-bed reactors. Catal. Today, 144 (2009), pp. 48–54.
Vohra, M.S., Malik, A.A., Al-suwaiyan, M.S. & Bukhari, A.A.: TiO2 -assisted photocatalytic removal of
phenol: effect of co-pollutants. International Journal of Applied Environmental Sciences, 4 (2009), pp.
33–45.
Vohra, A., Goswami, D.Y., Deshpande, D. A. & Block, S. S.: Enhanced photocatalytic inactivation of bacterial
spores on surfaces in air. J. Ind. Microbiol. Biotechnol. 32 (2005), pp. 364–370.
Wolfrum, E. J., Huang, J., Blake, D. M., Maness, P., Huang, Z. & Fiest, J.: Photocatalytic oxidation of bacteria,
bacterial and fungal spores, and model biofilm components to carbon dioxide on titanium dioxide coated
surfaces. Environ. Sci. Technol 36 (2002), pp. 3412–3419
Yamazaki-Nishida, S., Nagano, K. J., Phillips, L. A., Cervera-March, S. & Anderson, M. A.: Photocatalytic
degradation of trichloroethylene in the gas phase using TiO2 pellets. Journal of Photochemical and
Photobiology A: Chemical, 70 (1993), pp. 95–99.
Zhao, J. & Yang, X.: Photocatalytic oxidation for indoor air purification: A literature review. Building and
Environment, 38 (2003), pp. 645–654.
Zhao, J., Krishna, V., Hua, B., Moudgil, B. & Koopman, B.: Effect of UV-A irradiance on photocatalytic and
UV-A inactivation of Bacillus cereus spores. J. Photochem. Photobiol. B., 94 (2009), pp. 96–100.
196 S.M. Zacarías et al.

Zacarías, S. M., Vaccari, M. C., Irazoqui, H. A., Imoberdorf, G. E. & Alfano, O. M.: Effect of the radiation
flux on the photocatalytic inactivation of spores of Bacillus subtilis. J. Photochem. Photobiol. A, 214
(2010), pp. 171–180.
Zacarías, S. M., Satuf, M. L., Vaccari, M. C. & Alfano, O. M.: Efficiency evaluation of different TiO2 coat-
ings on the photocatalytic inactivation of airborne bacterial spores. Industrial & Engineering Chemistry
Research, 51 (2012), pp. 13599–13608.
CHAPTER 12

Water decontamination by heterogeneous photo-Fenton processes


over iron, iron minerals and iron-modified clays

Andrea De León, Marta Sergio, Juan Bussi, Guadalupe Ortiz de la Plata,


Alberto Cassano & Orlando Alfano

12.1 INTRODUCTION

The development of new clean technologies that reduce effluent pollution or mitigate the inef-
ficiency of conventional effluent treatment methods constitutes a significant contribution to the
fulfillment of sustainable development goals. Large-scale effluent treatment plants rely on a
number of physical and/or physicochemical methods, such as filtration, reverse osmosis and
activated carbon adsorption, chemical methods, such as chlorination, ozonization and thermal oxi-
dation, and biological methods. However, the removal of barely biodegradable and/or highly toxic
pollutants—i.e., harmful even in extremely small doses—by the currently used methods is often
highly inefficient and/or costly. Current laboratory- and pilot-scale studies are aimed at developing
non-conventional, more effective methods for the efficacious treatment of pesticides, heavy met-
als, chlorinated hydrocarbons, dyes disinfectants, and antibiotics, among other pharmaceutical
and biomedical contaminants.
Among the non-conventional methods, advanced oxidation techniques (AOTs) are based on the
high oxidizing power of hydroxyl radicals (HO• ) and their rapid, non-selective reaction capacity
(Ameta et al., 2012; Legrini et al., 1993). In view of the short half-life of hydroxyl radicals, they
are generated in situ by different oxidizing agents (H2 O2 , O3 , O2 ) that may be either present in or
expressly added to the reaction medium. AOTs enable the efficient conversion of organic matter
into stable inorganic compounds like water, carbon dioxide and salts (mineralization). AOTs are
particularly useful for destroying biologically toxic or non-degradable organic compounds, such
as aromatics, pesticides, petroleum constituents, and volatile species in waste water.
One of the most widely known AOTs is based on the Fenton system, in which the generation of
hydroxyl radicals results from reaction between H2 O2 and Fe2+ in aqueous solution at temperature
close to ambient conditions (Fenton, 1894; Safarzadeh-Amiri et al., 1996, Bautista et al., 2008),
according to:
Fe2+ + H2 O2 → Fe3+ + OH− + HO• (12.1)

Due to the low solubility of Fe3+ at pH > 3.2, the reaction (Eq. 12.1) is conducted at pH below
this value.
Fenton-based AOTs represent a low-cost alternative for the removal of natural organic matter
prior to the disinfection of water supplies aimed for human consumption, in view of the mild
working conditions required, as well as the low cost of both hydrogen peroxide and iron (the
second most abundant metal and the fourth most abundant element on earth) and their associated
high efficiency in the process.
The irradiation of the solution with an appropriate light source (λ < 400 nm) contributes to the
generation of radicals and the regeneration of Fe2+ by photolysis of the Fe(OH)2+ complex:

Fe(OH)2+ + hν → Fe2+ + HO• (λ < 400 nm) (12.2)


197
198 A. De León et al.

The thus-called photo-Fenton system reaches the maximum reaction rate at pH close to 3, a
value corresponding to the maximum concentration of the complex ion Fe(OH)2+ (Pignatello,
1992; Bolton, 2001).
Although the system is best characterized by reactions (11.1) and (11.2), other reactions also
occur in varying degrees, according to experimental conditions (light wavelength and intensity,
temperature, relative proportions of oxidizing agent, catalyst and organic matter), thus, affecting
either positively or negatively the overall process efficiency. In particular, the negative effect of
a high H2 O2 concentration is ascribed to the HO• radical scavenging reaction (Bautista et al.,
2008; Pignatello, 1992, Pignatello et al., 1999) according to:

H2 O2 + HO• → HO•2 + H2 O (12.3)

The potential low cost of solar light as main irradiation source (4–5% of the solar spectrum is
comprised within the UV region between 300 nm and 400 nm) is driving interest in photo-Fenton
based AOTs.
In the photo-Fenton system, iron acts as a homogeneous catalyst, and the iron concentrations
necessary to attain reasonable efficiency levels have been reported to exceed the generally agreed/
limits for effluent discharges into watercourses. As a result, iron recovery by precipitation at pH
above 4 is required, leading to the generation of sludge associated with environmental risk even
after final disposal. An interesting alternative is to dissolve the generated sludge with a view to
recovering iron for later re-use. These options require manpower and, in the latter case, chemical
reagents that significantly increase the overall process cost.
The disadvantages associated with the use of Fe3+ in solution (homogeneous photo-Fenton
process) may be avoided, as in other catalytic processes, using iron compounds in a solid phase,
thus, facilitating the separation and recovery of iron for later re-use. This leads to the thus-called
heterogeneous photo-Fenton processes, in which the resulting catalytic activity is attributed to
iron contained in the solid phase, although part of it may be solubilized and act as a catalyst in
solution. An additional advantage reported for the heterogeneous system results from a wider
range of working pH than that associated with the homogeneous process.
Iron minerals or iron species supported on different solids have been tested for their catalytic
performance in heterogeneous photo-Fenton processes applied to the degradation of a wide variety
of contaminants. These have included mainly dyes and pigments (Hou et al., 2011) as well
as phenolic compounds (Zhang et al., 2011), in addition to a variety of pharmaceutical drugs
(Ayodele et al., 2012) and agro-chemicals (Bouras et al., 2007) representative of varied industrial
and municipal effluents (Galeano et al., 2011).
Interest has been placed on the stability of these catalysts, which enables their re-use in multiple
cycles. The reutilization of certain catalysts in multiple processes has been reported (Sum et al.,
2005), although, in other works, iron leaching has been demonstrated to compromise catalyst
stability (Ortiz de la Plata et al., 2010a; Son et al., 2009). The loss of iron has been attributed to
diverse causes, such as the formation of highly soluble iron-reaction intermediate complexes (e.g.,
oxalate), and the thus required low pH (normally pH 3), high H2 O2 /reactant ratio, and temperature
and radiation intensity conditions (Najjar et al., 2007). None the less, the concentration of iron in
solution does not compromise the water quality in the treatment outlet. In some cases, the amount
of solubilized iron decreases in advanced stages of the mineralization process, a fact that has been
ascribed to the destruction of organic intermediates that maintained iron in solution and to the
fixation of iron ions onto the solid catalyst (Chen and Zhu, 2011; Feng et al., 2009; Sum et al.,
2005).
In view of the environmental impact of AOTs, particularly heterogeneous photo-Fenton pro-
cesses, the development of new solids exhibiting catalytic activity remains an issue of academic
as well as applied interest.
This chapter presents the results of a number of catalyst development and performance studies
focused on the use of iron-based catalysts in heterogeneous photo-Fenton processes. Special
Water decontamination by heterogeneous photo-Fenton processes 199

emphasis is placed on the influence of a set of parameters relevant to the catalyst preparation
process and the catalytic reaction process under photo-Fenton conditions.

12.2 CATALYSTS FOR USE IN HETEROGENEOUS PHOTO-FENTON PROCESSES

12.2.1 Iron and iron minerals


Among the iron compounds that have been widely studied as catalysts in heterogeneous photo-
Fenton systems are the iron oxides and hydroxides, like hematite and goethite, in a wide variety of
forms and origins, from amorphous to crystalline, both natural or synthetic (Huang and Huang,
2009; Ortiz de la Plata et al., 2010b) and zerovalent iron (Ortiz de la Plata et al., 2012; Son et al.,
2009).
Goethite (α-FeOOH) is a natively found iron oxyhydroxide, has been extensively studied since
it combines attractive properties for large-scale applications; standing out (i) thermodynamic
stability, (ii) controllable leaching of iron into the solution and (iii) the possibility of using solar
radiation as a primary source of energy for a photo-Fenton alternative (Ortiz de la Plata et al.,
2008).
Most recently have appeared various studies that employ zerovalent iron (ZVI) as iron source
in the photo-Fenton process. There is a wide range of ZVI solids studied, varying in source and
granulometry. A fraction of these materials, the nanoparticulated ZVI, (Morgada et al., 2009; Xu
et al., 2011; Ortiz de la Plata et al., 2012) possess also the possibility of reduce organic components
in the appropriate conditions, so can be used in advanced oxidation Fenton technologies, or as
reductive agents, combined or not with UV radiation. Nano ZVI has been widely studied due to
its high reactivity, based fundamentally on its high surface contact area, and low leaching.

12.2.2 Supported and immobilized iron species


Various solids have been used to support or immobilize iron species based on different commonly
used catalyst preparation procedures. Among the solids used as support are organic polymers like
Nafion—in which Na+ in the -SO3 Na group is substituted by Fe3+ (Fernández et al., 1999)—,
synthetic clays like laponite (Feng et al., 2003a; 2003b; Iurascu et al., 2009), natural clays like
bentonite, beidellite, nontronite (Catrinescu et al., 2011; Cheng et al., 2008; Feng et al., 2004a)
clays modified by procedures like acid activation and/or pillaring (Ayodele et al., 2012; Najjar
et al., 2007; Zhang et al., 2011), porous materials like resins (Liou et al., 2005) and zeolites
(Noorjahan et al., 2005; Martínez et al., 2005), among other materials, such as silica, alumina
and/or carbons (Ramirez et al., 2007; Rodríguez et al., 2010), and micas like vermiculite (Chen
et al., 2010).

12.2.3 Iron species supported on clays


The use of clays as catalysts has been known since the start of the development of the petro-
chemical industry. Clays are layered aluminosilicates with physicochemical properties that make
them suitable for a wide range of applications. In particular, clays of the smectite group—either
in their natural form or modified by varied procedures—have been widely used on account of
their small particle size, high cation-exchange capacity and swelling capacity (increase in the
interlayer spacing leading to the generation of a volume with the potential for adsorption and
catalysis). In addition to the original treatments—entailing only acid activation—a new set of
procedures are currently available. Among these, pillaring is based on the substitution of ions
that are exchangeable with bulky cations, which are later converted into oxides by calcination.
These oxides form pillars that prop the clay layers apart, thus increasing the exposed surface area,
i.e., effectively available for adsorption and catalysis. The resulting solids, referred to as PILCs
(Pillared Interlayered Clays), are thermally stable and have a high specific surface area, and their
catalytic activity has been associated with the mineral, the pillar, and even a synergistic effect
between both. The most widely studied polycations are those containing Al, Ti, Zr, Fe, Cr, or a
200 A. De León et al.

combination of two metals. Numerous studies on the preparation, characterization and application
of such catalysts have been compiled in bibliographic reviews (Centi and Perathoner, 2008;
Embaid et al., 2011; Gil and Gandía, 2000; Kloprogge, 1998).
In a recent review, Herney-Ramirez et al. (2010) reported the excellent properties of PILCs in
photo-Fenton processes applied to the treatment of contaminated water and analyzed the influence
of different experimental conditions (light wavelength and intensity, initial amount of H2 O2 ,
catalyst load, pH and temperature). In addition, the authors describe different PILC preparation
procedures based on the use of iron (Fe-PILC)—and other metals like copper—and address aspects
of applied interest, such as their stability and use under visible light. Different particularly relevant
contaminants have been treated with Fe-PILCs. Among these are phenolic compounds and dyes,
which are widely used in different industries (e.g., paper and textile).

12.3 EXPERIMENTAL

In this section, the most relevant aspects defining the conditions under which the experimental
studies were conducted are described.

12.3.1 Catalysts
The here-reported results were obtained using iron pillared clays prepared in the laboratory and
commercial goethite and nanoparticulate zerovalent iron (nZVI).

12.3.1.1 Fe-PILCs
Conventional PILC preparation methods rely on the use of a clay fraction (aggregate size smaller
than 2 µm) requiring costly separation processes associated with high water and time inputs. As
the clay used in our studies is a montmorillonite (Bañado de Medina, Uruguay), containing not
more than 10% of impurities (only silica and quartz), a less time and labor consuming preparation
method was designed. Different size fractions were selected by sieving the clay, which was used
without pre-treatment other than drying and milling.
The clay fractions were ion exchanged with the iron complex [Fe3 (OCOCH3 )7 OH.2H2 O]NO3 ,
which was prepared according to Yamanaka et al. (1984). A solution containing the complex
was added to a clay suspension and the thus resulting suspension aged at room temperature.
The exchanged clay was separated by filtration and washed repeatedly with deionized water,
stove-dried at 60◦ C and calcined in air (De León et al., 2008; Noya et al., 2011).
Different Fe-PILC catalysts were obtained using different fractions of the clay (differing in
aggregate size) and calcination temperatures, as indicated below.

12.3.1.2 Goethite
Goethite – α-FeOOH – (particulate, catalytic grade) was purchased from Aldrich. Particles in the
range from 75 to 150 µm were selected.

12.3.1.3 Zerovalent iron


Zerovalent iron nanoparticles (nZVI) were provided by Nanotek S.A. (www.nanotek-sa.com.ar).
The product is a dark suspension of nanoparticles with magnetic properties (Ortiz de la Plata
et al., 2012).

12.3.2 Catalyst characterization


Thermogravimetric analysis (TGA) and differential thermal analysis (DTA) of the exchanged
clays (prior to calcination) enabled the determination of the appropriate calcination temperature.
Catalyst textural properties were determined based on nitrogen adsorption-desorption isotherms
performed at −196◦ C. Scanning electron microscopy (SEM) micrographs of the clay and Fe-PILC
samples were obtained to observe the morphology of these materials. Fe-PILC was acid treated
Water decontamination by heterogeneous photo-Fenton processes 201

(HCl) and iron in the obtained solution was quantified by atomic absorption to determine the iron
content in the catalyst.

12.3.3 Photocatalytic tests


Aqueous solutions of two dyes and one chlorinated phenolic compound: methylene blue (MB)
and methyl orange (MO) and 2-chlorophenol (2-CP), respectively, were used for degradation tests
under photo-Fenton conditions.
Two different batch reactors, according to the following main characteristics, were used.

12.3.3.1 Fluidized bed batch reactor


This reactor was used in dye degradation tests. It was constructed of a borosilicate glass tube
with a sintered glass stopper on its lower end to allow the influx of compressed air necessary to
maintain the catalyst in suspension and in intimate contact with the reaction mixture. An air flow
rate of 1.5 L min−1 was used. Radiation was supplied by four Phillips TL’D 18/08 lamps (emission
range of 340 nm through 420 nm, spectrum maximum at 360 nm) symmetrically disposed around
the reactor. The temperature was controlled by means of thermostatted water circulating through
a U-tube located in the reactor center. The initial time was considered to be the moment of catalyst
addition.

12.3.3.2 Stirred batch reactor


It was employed in the 2-chlorophenol degradation studies. It consists in a cylindrical, well-
stirred, batch reactor irradiated through a transparent radiation bottom. The reactor was also
equipped with internal glass heat exchangers, connected to a thermostatic bath for controlling
temperature, and an external insulation made of K-Wool. The described device allows irradiating
the suspension with photons in the wavelength range corresponding to UVA (specifically 340
to 380 nm) with a TL’D 18/08 Phillip’s black light lamp placed at the focal axis of a parabolic
reflector. 2.0 L reaction volume was used, which implies an optical path of approximately 9 cm.

12.3.4 Analytical techniques


The colorants (MB and MO) were quantified spectrophotometrically with a Shimadzu UV 1203
equipment. 2-CP was quantified by HPLC employing a Waters UV detector spectrophotometer.
Total organic carbon (TOC) was measured with a Shimadzu TOC-5000 Analyzer. H2 O2 was mea-
sured spectrophotometrically with a modified iodometric technique; ferrous ions were analyzed
spectrophotometrically by means of absorbance of the Fe2+ -phenanthroline complex at 510 nm
(total dissolved iron was reduced employing ascorbic acid in 0.45 µm filtered samples). A Cary
100 Bio spectrophotometer was used for both analyses.

12.4 CATALYTIC ACTIVITY

In this section, different results obtained in our laboratory are presented with the aim of sys-
tematizing a number of findings that contribute to an improved understanding of these types of
photocatalytic systems.

12.4.1 Iron-pillared clays used for dye degradation


12.4.1.1 Contribution of different processes entailed in contaminant removal
The evaluation of the relative influence of those parameters involved in the photo-Fenton reaction
contributes to an improved understanding of the overall process and to a greater degree of control
over the reacting system. In fact, both catalytic mechanisms (Fenton, photo-Fenton) and non-
catalytic ones (adsorption, photolysis) contribute to the degradation of contaminants under typical
experimental photo-Fenton conditions. The results of a series of tests designed to assess the
relevance of different reaction parameters are presented in this section.
202 A. De León et al.

Figure 12.1. MB remaining in the solution as a function of time. Catalyst load: 1.0 g L−1 , [MB] = 0.2 mM,
[H2 O2 ] = 10 mM. Adapted from Catalysis Today 133–135, M.A. De León, J. Castiglioni,
J. Bussi, M. Sergio, Catalytic activity of an iron-pillared montmorillonitic clay mineral in
heterogeneous photo-Fenton process, 600–605, Copyright (2008), with permission from
Elsevier.

Figure 12.1 shows the results of MB removal tests under different experimental conditions using
1.0 g L−1 of the catalyst obtained by iron pillaring the clay fraction containing aggregates of size
smaller than 250 µm (Fe-PILC), followed by calcination at 400◦ C. A rapid decrease was observed
in the MB concentration in solution regardless of the experimental conditions, reaching similar
removal levels after 20 minutes. The above suggests that the contribution of adsorption to the
discoloration of the solution was the highest among the phenomena involved. At greater reaction
times, the different removal curves depart from each other, reflecting the different contribution
of each phenomenon.
The adsorption curve (MB, Fe-PILC) shows a rapid decrease in the amount of MB remaining in
solution within the first 20 minutes of contact and a decrease in the adsorption rate (curve slope)
as the equilibrium is reached at a removal efficiency of 63% after 180 minutes. A considerable
degree of adsorption was expected in view of the high affinity of different solid surfaces for
the organic, positively charged MB ion in aqueous solution (Hang and Brindley, 1970). Indeed,
the cation-exchange capacity retained by PILCs (Martin-Luengo et al., 1989) appears to further
improve their affinity for MB.
In the Fenton test (MB, Fe-PILC, H2 O2 , pH 3) the MB removal efficiency was as high as 80%,
i.e., higher than that attained by adsorption, though significantly lower than the 93% attained in
the photo-Fenton test (MB, Fe-PILC, H2 O2 , pH 3, UV). These results show that both the Fenton
and photo-Fenton processes take place using the tested Fe-PILC, promoting the production of HO•
radicals as a result of the decomposition of hydrogen peroxide. Lin and Gurol (1998) proposed
that the catalytic decomposition of hydrogen peroxide over iron-containing solid catalysts occurs
through the formation of complex peroxidic species on Fe(III) active sites on the catalyst surface.
≡Fe(III)OH + H2 O2 → ≡Fe(II)OH2 + HO•2 (12.4)

A series of reactions involving the transfer of electrons generate active Fe(II) species and
hydroperoxyl radicals (HO•2 ). Reduced iron sites are capable of reacting either with hydrogen
peroxide or with oxygen, thus reoxidizing the iron. The additional action of UV radiation in the
heterogeneous photo-Fenton system accelerates the formation of HO• radicals, possibly by photo-
reduction of Fe(III) active sites in like manner to that proposed for the homogeneous process.
Later reaction of MB, adsorbed on the surface of the Fe-PILC, with HO• radicals leads to the
Water decontamination by heterogeneous photo-Fenton processes 203

Figure 12.2. SEM micrographs for (a) clay and (b) Fe-PILC250–450 .

formation of reaction intermediates and eventually CO2 and H2 O as final products. The adsorption
of reactants on the catalyst surface appears to be a condition for the reaction to occur.
The curve obtained in the discoloration test in which H2 O2 was not used (MB, Fe-PILC, pH
3, UV) practically coincides with that obtained under Fenton conditions. In the absence of H2 O2 ,
the formation of hydroxyl radicals may occur directly from hydroxyl complexes of photo-active
Fe(III) on the Fe-PILC surface according to reaction (12.2) proposed for the homogeneous system.
The reoxidation of Fe(II) is driven by the action of oxygen according to the following reactions
(Lin and Gurol, 1998):
Fe(II) + O2 + H2 O → Fe(III)(OH)2+ + HO•2 (12.5)
2HO•2 → H2 O2 + O2 (12.6)

Hydroxyl radicals may also be generated by photo-activation mechanisms involving the for-
mation of electron-hole pairs (e− -h+ ) in semiconducting metal oxides like TiO2 , ZnO and Fe2 O3
(Bandara et al., 2001). Feng reported the presence of tetragonal Fe2 O3 in iron-pillared clays (Feng
et al., 2003b; 2004b). Also FeOOH species have been identified in iron-pillared clays prepared
by a similar method to that of this study (Martin-Luengo et al., 1989), and their catalytic activity
proposed (Wang et al., 2009).
Whereas these Fe-PILCs have been proved to lead to a number of phenomena that contribute
to the removal of the tested colorant (MB) in different degrees, the greatest contribution has been
observed to occur under photo-Fenton conditions.

12.4.1.2 Influence of the clay aggregate size used for Fe-PILC preparation
The size of clay aggregates subject to pillaring has an influence on the properties of the final
catalyst (textural properties, amount of incorporated cation), which may ultimately affect its cat-
alytic performance. The results of MB removal tests in aqueous solution obtained using different
Fe-PILCs are shown in this section.
Catalysts were prepared by pillaring from two clay fractions of aggregate size under 250 µm
(Fe-PILC<250 ) and between 250 µm and 450 µm, (Fe-PILC250–450 ), respectively, followed by
calcination at 400◦ C. The iron content of the resulting catalysts was 21% for Fe-PILC<250 and
14% for Fe-PILC250–450 .
The morphology of the host clay and the Fe-PILC250–450 can be appreciated in the SEM micro-
graphs shown in Figure 12.2. The well-ordered structure characteristic for the flat particles of the
montmorillonite is confirmed by Figure 12.2a in which interparticle pores can only be observed.
The micrograph in Figure 12.2b evidences that the pillaring process has preserved the lamellar
structure of the montmorillonite; however, this process caused a little swelling and resulted in a
less ordered structure.
Figure 12.3 shows the amount of MB remaining in solution in adsorption tests under photo-
Fenton conditions. The adsorption curves obtained for the two catalysts follow the above-described
204 A. De León et al.

Figure 12.3. MB remaining in the solution (%) as a function of time for the catalysts Fe-PILC<250 and Fe-
PILC250–450 . Catalyst load: 1 g L−1 . Adsorption: MB, Fe-PILC; Photo-Fenton: MB, Fe-PILC,
H2 O2 , pH 3, UV; [MB] = 0.2 mM, [H2 O2 ] = 10 mM. Reprinted from Catalysis Today 133–
135, M.A. De León, J. Castiglioni, J. Bussi, M. Sergio, Catalytic activity of an iron-pillared
montmorillonitic clay mineral in heterogeneous photo-Fenton process, 600–605, Copyright
(2008), with permission from Elsevier.

140

120

100 Fe-PILC 250–450

80

60

40

20

0
0.0 0.2 0.4 0.6 0.8 1.0
Relative pressure (p/p˚)

Figure 12.4. Nitrogen adsorption-desorption isotherms at −196◦ C for Fe-PILC < 250 and Fe-PILC250-
450; Full symbols: adsorption; Open symbols: desorption. Reprinted from Catalysis Today
133–135, M.A. De León, J. Castiglioni, J. Bussi, M. Sergio, Catalytic activity of an iron-
pillared montmorillonitic clay mineral in heterogeneous photo-Fenton process, 600–605,
Copyright (2008), with permission from Elsevier.

behavior, although the final MB amount in solution was found to differ, with 48 and 37% for
Fe-PILC250–450 and Fe-PILC<250 , respectively.
Figure 12.4 shows the nitrogen adsorption-desorption isotherms at −196◦ C for the two cat-
alysts. Both curves have the characteristic shape of those obtained with pillared clays. A high
nitrogen adsorption at low relative pressures is indicative of a large adsorption volume generated
at the micropore level. The gradual increase in the rate of nitrogen adsorption (though in a lesser
Water decontamination by heterogeneous photo-Fenton processes 205

proportion) in the region of intermediate relative pressures, through which mesopores are filled,
is ascribed to adsorption occurring within gaps between platelets in the solid, consistently with
the observed small H4-type hysteresis loop with parallel branches closing at relative pressures in
the proximity of 0.4 (Rouquerol et al., 1999).
The difference in the amount of MB removed by each catalyst may be correlated with the
specific surface area of the solids, determined from nitrogen adsorption data. In effect, the solid
obtained from the smaller-sized aggregates of the clay, Fe-PILC<250 , with a specific surface area
of 212 m2 g−1 , retains 63% of the starting MB in solution by adsorption, while Fe-PILC250–450 ,
with a specific surface area of 140 m2 g−1 , retains 52% of MB.
The curves obtained in photo-Fenton tests (full lines in Fig. 12.3) show the combined effects of
catalyst, light and H2 O2 in the discoloration of the MB solution when the pH was initially adjusted
to 3. A rapid decrease in the MB concentration was observed within the first 20 minutes of the
time of contact. A significant overlap between the adsorption and photo-Fenton curves suggests
the absence of a contribution resulting from the action of light or hydrogen peroxide over this
period. The reduction in dye concentration is sharper for the catalyst obtained from the smaller
aggregate size fraction (Fe-PILC<250 ), having a higher specific surface area. Using this catalyst,
the contribution of the photo-catalytic process becomes evident after 20 minutes of reaction, as
reflected in a higher curve slope (removal rate) than that of the corresponding adsorption curve,
attaining practically full MB removal, 93%, over the time of the test (180 minutes).
Using the catalyst Fe-PILC250−450 , the percentage of MB removal between 20 and 80 minutes
of reaction is lower than that attained by adsorption. However, this situation is reverted and
at 180 minutes MB removal reached 87%, a value similar to that found for Fe-PILC<250 . An
apparent inflexion point in this curve—not observed in the Fe-PILC<250 curve—may be attributed
to different properties of the catalyst. Analytical determination of the iron content shows that
the catalyst prepared with the larger aggregate size fraction (Fe-PILC250−450 ) contains 45% less
iron than that prepared with the smaller aggregate size fraction (Fe-PILC<250 ), resulting in a
smaller number of active sites and therefore, a reduced HO• generation rate. Besides, a greater
aggregate size results in a slower access of reactants to the active sites—in particular for MB, in
view of its large molecular size—, resulting in a higher H2 O2 concentration within the porous
catalyst structure, thus, favoring the scavenging of HO• radicals over an initial stage, according
to reaction (3).
The influence of the hydroxyl radical scavenging reaction is assumed to decrease in more
advanced stages (concavity change at 120 minutes) on account of a lower concentration of H2 O2 ,
thus favoring the action of HO• radicals on MB. In this regard, unpublished results showed a
marked increase in the initial rate of MB removal with decreasing H2 O2 concentration for this
catalyst. In contrast, no such point of inflexion was observed in tests using the Fe-PILC<250
catalyst; neither did the concentration of H2 O2 appear to have an influence on the MB removal
efficiency. All the above is consistent with the smaller amount of iron incorporated through
pillaring by the Fe-PILC250−400 catalyst and the fact that a lower degree of exposure—associated
with a larger aggregate size—unfavors the accessibility of reactants, especially for MB.
In sum, whereas the size of the clay aggregates used for pillaring does not appreciably affect
the capacity of the catalyst to decolorize a MB solution, it does have an influence on the profile
of removal curves as a function of time, an influence which may be ascribed to the effects of
competition and/or catalyst textural properties.

12.4.1.3 Influence of the initial pH of the reaction medium


As mentioned in the Introduction, the effectiveness of homogeneous photo-Fenton processes is
closely associated with the pH of the reaction medium. The formation of highly oxidizing HO•
radicals results from the photolysis of Fe(OH)2+ species, which are stable at pH values in the prox-
imity of 3. Fe(III) species are converted into Fe+3 (H2 O)6 at lower pH and into Fe(OH)3 at higher
pH values, the latter two species being less active in the formation of HO• radicals (Safarzadeh-
Amiri et al., 1996). A narrow pH range over which the homogeneous process is efficient appears
as an inconvenience, in particular when natural effluents are concerned, as the control of medium
206 A. De León et al.

Figure 12.5. Influence of pH on MB (%) remaining in solution as a function of time in photo-Fenton


tests. Catalyst load: 1 g L−1 . [MB] = 0.2 mM. [H2 O2 ] = 10 mM. Initial pH was adjusted to
the required value with diluted HCl. Adsorption curve is shown for comparison. Adapted from
Catalysis Today 133–135, M.A. De León, J. Castiglioni, J. Bussi, M. Sergio, Catalytic activity
of an iron-pillared montmorillonitic clay mineral in heterogeneous photo-Fenton process,
600–605, Copyright (2008), with permission from Elsevier.

acidity becomes a necessity. Therefore, an additional advantage of heterogeneous processes may


be associated with its efficiency over a wider range of pH.
Below are presented the results of one study on the influence of pH on the removal efficiency
by heterogeneous photo-Fenton processes using iron-pillared clay catalysts.
Figure 12.5 shows the influence of pH on the amount of MB remaining in the solution under
photo-Fenton conditions using a Fe-PILC obtained from the clay fraction of aggregate size smaller
than 250 µm by pillaring and calcination at 400◦ C. There is strong overlap among the curves
obtained at different pH values, suggesting a practically negligible influence of pH on MB removal
over the range of 3 to 6. These results suggest that the influence of pH may be attenuated in
heterogeneous systems, presumably due to the anchoring of Fe(III) species—responsible for the
formation of HO• radicals by photo-reduction—in the host montmorillonite clay. Moreover, the
small volume containing active species within the clay’s interlayer spacing may contribute to
buffering the effects of medium pH.
Figure 12.6 shows the effect of pH on the amount of MO remaining in solution using Fe-
PILC (obtained by pillaring the clay fraction of aggregate size between 250 µm and 450 µm and
calcination at 350◦ C) under photo-Fenton conditions.
Whereas practically complete dye removal (91%) was attained at pH 3, at pH 6, the initial
concentration of MO in solution remained practically constant even after 240 minutes.
The marked effect of pH on the removal of MO differs from the results obtained in MB
removal tests, showing the strong dependence of the heterogeneous photo-Fenton process on the
characteristics of the molecule being degraded, which may be ascribed to the type of interaction
between the latter and the Fe-PILC. The pillared clay has been reported to retain a certain degree
of cation-exchange capacity (Martin-Luengo et al., 1989). This result in a residual negative charge
on the clay platelets, constituting adsorption sites especially suitable for positively charged species
(De León et al., 2008). A pKa of 3.7 of MO shows that the polarity of the molecule changes over
the studied pH range (3 through 6). Figure 12.7 shows the MO molecular structure at different
pH. At pH 6 (pH > pKa ), MO is negatively charged and its adsorption on the Fe-PILC surface is
hindered, resulting in ineffective catalysis for degradation. However, in the range of acid pH below
Water decontamination by heterogeneous photo-Fenton processes 207

Figure 12.6. Influence of pH on methyl orange (%) remaining in solution as a function of time in
photo-Fenton tests. Catalyst load: 1 g L−1 . [MO] = 0.2 mM. [H2 O2 ] = 10 mM. Initial pH was
adjusted to the required value with diluted HCl. Adsorption curve is shown for comparison.
Adapted from Noya et al. (2011).

Figure 12.7. Molecular structure of methyl orange, pKa = 3.7: (a) pH > pKa ; (b) pH < pKa .

Figure 12.8. Molecular structure of methylene blue.

the pK a the predominant species is the zwitterion (Fig. 12.7b), which promotes dye adsorption
on the Fe-PILC surface and its photocatalytic degradation.
In contrast, the fact that the MB molecule represented in Figure 12.8 is positively charged irre-
spective of pH—over the range studied—is in agreement with the apparent lack of pH dependence
of the efficiency of MB catalytic removal.
These results suggest the important role of dye adsorption as initial step in the mechanism that
leads to dye degradation by radicals HO• (see reaction (12.7)). The degradation of the adsorbed
and /or photosensitized dye could be favored by direct electronic exchange with Fe (III) sites by
mechanisms similar to those proposed in the literature (Chen et al., 2009; Cheng et al., 2008):
≡Fe(III) + Dye(Dye∗ ) → ≡Fe(II) + Dye+• (12.7)

The reduced iron sites contribute to increase HO• production by the typical Fenton
reaction (12.1).
208 A. De León et al.

Figure 12.9. DrTGA profile for the starting clay and the ion-exchanged clay.

These results highlight the contribution of the initial step of substrate adsorption in the colorant
degradation mechanism involving HO• radicals.
Dye adsorption, as a step prior to chemical degradation, confirms the heterogeneous catalysis
hypothesis as the main degradation route, based on the studied dye. The scarce influence of
pH on the MB removal efficiency suggests that catalytically active species may not have been
generated by the partial dissolution of iron, as the here-reported results are inconsistent with an
otherwise expected marked dependence on pH, i.e., typically associated with the homogeneous
photo-Fenton process.

12.4.1.4 Selection of the temperature for calcination of the exchanged clay


The PILC preparation process entails the calcination of the exchanged-clay in order to promote
the formation of iron oxide pillars that generate the microporous structure typical of PILCs. It
has been proposed that the pillars containing the active species are highly dispersed within the
interlayer space of pillared clays, thereby favoring its atomic efficiency during catalysis.
In this study, the temperature for the calcination of the ion-exchanged solids was selected
based on the results of thermogravimetric analysis (TGA). Figure 12.9 shows the derivative
thermogravimetric analysis (DrTGA) profile of the starting clay and the ion-exchanged clay. The
DrTGA profile of the clay shows a peak at about 100◦ C corresponding to a mass decrease ascribed
to the loss of weakly-bound water within the clay structure, amounting to 7.9% of the initial mass.
A second mass loss is evidenced by a peak in the curve situated in the range of 340 to 470◦ C,
showing a minimum at 420◦ C attributed to the transformation of Fe(III) compounds present in
the host clay (Hatakeyma and Liu, 1998). A further peak observed at around 660◦ C is ascribed
to the dehydroxylation of the clay.
The DrTGA for the exchanged clay shows the loss of weakly-bound water within the structure
(peak at around 100◦ C) amounts to 5.9% of the initial mass. A second peak in the DrTGA within
the range of 220 to 270◦ C, not observed in the clay profile, is ascribed to the decomposition of
the iron complex incorporated during the ion-exchange step, leading to the formation of pillars
between clay layers (Martin-Luengo et al., 1989; Zhang et al., 2011). An additional peak is
observed in the same temperature range as for the clay (340◦ C to 470◦ C), though the minimum
occurs at slightly lower values (405◦ C), possibly due to pillar-clay interactions.
According to these results, a calcination temperature higher than 270◦ C would result in a solid
with the typical structure of PILCs. Calcination temperatures of 350 and 500◦ C were selected
for this study, the former value ensuring the formation of iron oxide pillars and the latter the
stabilization of the host clay structure. The resulting catalysts are referred to as Fe-PILC-350 and
Fe-PILC-500, respectively.
Water decontamination by heterogeneous photo-Fenton processes 209

100

80

60

40
Fe-PILC-350

20 Fe-PILC-500

0
0.0 0.2 0.4 0.6 0.8 1.0
(p/p˚)

Figure 12.10. Nitrogen adsorption-desorption isotherms at −196◦ C. Adapted from Noya et al. (2011).

Table 12.1. Textural parameters of Fe-PILCs and the host clay derived from
nitrogen adsorption data at −196◦ C.Adapted from Noya et al. (2011).

Clay Fe-PILC-350 Fe-PILC-500

SBET (m2 g−1 ) 29 139 130


Vµpore (cm3 g−1 ) 0.011 0.059 0.054
VTotal (cm3 g−1 ) 0.050 0.115 0.114

Figure 12.10 shows the nitrogen adsorption isotherms at −196◦ C for the two catalysts and the
starting clay. As discussed in 12.4.1.2 above, the shape of these isotherms confirms the formation
of the microporous structure characteristic of PILCs. Textural parameters of the host clay and the
Fe-PILC obtained by calcination at different temperatures are presented in Table 12.1.
It was found that the effect of pillaring has led to a 4.5-fold increase in the specific surface
area and micropore volume of the starting clay. The specific micropore volume accounts for 50%
of the specific total pore volume of the Fe-PILCs (determined from adsorption data at a relative
pressure of 0.95), highlighting the microporous nature of these solids.
Figure 12.11 shows the results of adsorption and photo-Fenton tests at pH 3 using the catalysts
Fe-PILC-350 and Fe-PILC-500. Both catalysts led to a MO adsorption equilibrium with 25% of
MO remaining in solution, suggesting a similar MO adsorption capacity of the solids. This is in
line with a similar specific surface area and specific pore volume also found for the two solids
(Table 12.1). Nonetheless, the time required to attain the equilibrium differs for the two catalysts:
60 minutes for Fe-PILC-350 and 100 minutes for Fe-PILC-500, a fact that may be attributed
to greater diffusion constraints for MO within the microporous structure of Fe-PILC-500. This
indicates the effect of temperature-driven changes in the textural properties of the resulting Fe-
PILC that cannot be evidenced by nitrogen adsorption data, the latter being obtained in equilibrium
using a smaller molecule in gaseous phase.
The curves obtained under photo-Fenton conditions at pH 3 show that, although, at shorter
times, the dye removal level is lower than that attained by adsorption, at greater time values, it
exceeds the values corresponding to the adsorption equilibrium, reaching practically complete
removal after 4 hours. The above evidences the combined action of H2 O2 and radiation, which,
in the presence of iron contained in the catalyst, generates HO• radicals (Feng et al., 2003a;
2003b), responsible for the degradation of colorant adsorbed on the solid surface. A lower removal
210 A. De León et al.

Figure 12.11. MO (%) remaining in aqueous solution during adsorption and photo-Fenton tests using
Fe-PILC-350 and Fe-PILC-500. Adapted from Noya et al. (2011).

rate observed in photo-Fenton tests at intermediate time values may be attributed to diffusion
constraints and competition of H2 O2 for the catalyst active sites, resulting in a reduced rate of
colorant adsorption, as discussed in 12.4.1.2.

12.4.2 Fe-PILC, goethite and zerovalent iron in 2-chlorophenol degradation


In the present section, the behavior in the degradation of a model phenolic compound with different
solids (Fe-PILCs, goethite and nZVI) in heterogeneous photo-Fenton conditions is reported. The
model compound is the 2-chlorophenol (2-CP), and was selected because is representative of a
group of pollutants that are reticent to be degraded by conventional techniques. This substance
is toxic and widely employed in the industry (as disinfectant and germicide, and as a precursor
in the manufacture of pesticides and other chlorophenols) making it an important source of
environmental pollution.
The advance in reaction is monitored following 2-CP concentration (C2−CP ), H2 O2 concentra-
tion (Cp ), TOC concentration (CTOC ) and dissolved iron. The first three parameters are expressed
as relative values referred to each initial value, but not for the Fe in solution, which is reported in
the measured concentrations, usually mg L−1 .
Two Fe-PILCs were employed, with different iron content (w/w) of 17.6% and 6.1% named
Fe-PILC18 and Fe-PILC6 respectively. Figure 12.12 shows the results obtained in an assay using
the photo-Fenton conditions with the Fe-PILC18 with 1.0 g L−1 catalyst load. In the initial phase
a lag time is observed, with a really low 2-CP removal, followed by a marked increase in the
degradation rate after 60 minutes, reaching complete removal at 120 minutes. This removal is
accompanied by a reduction of the TOC that in the end of the experience represents a 15% of the
initial value, which shows the existence of reaction intermediates, among which Cl-benzoquinone
(ClBQ) is usually mentioned (Ortiz de la Plata et al., 2010b). It also can be seen an increase in the
concentration of dissolved Fe in the reaction medium, explainable by the solid leaching, probably
by reaction intermediates (such as ClBQ and certain carboxylic acids). In the initial phase of
the experiment Fe concentration increases very slowly, reaching values close to 0.2 mg L−1 at 60
minutes, increasing more rapidly to reach 3.1 mg L−1 at 120 minutes. It is interesting to point out
that the increase in the leaching rate accompanies the increase in the removal of 2-CP. Similar
behavior was observed with the Fe-PILC6 (Fig. 12.13), but with significantly lower values for the
final removal of 2-CP and the amount of dissolved Fe, 24% and 0.5 mg L−1 , respectively.
Figure 12.14 presents the obtained results employing two different goethite loads (0.5 and
2.0 g L−1 ). Evolution profiles of 2-CP concentration observed are similar to those obtained with
Water decontamination by heterogeneous photo-Fenton processes 211

1.0
3

0.8
2-CP
0.6 2
C/Co TOC
H2O2
0.4 Fe
1
0.2

0.0 0
0 30 60 90 120
Time (min)

Figure 12.12. Dimensionless concentration of reactants, products and iron vs. time for Fe-PILC18 . Catalyst
load = 1.0 g L−1 , C2-CP = 50 mg L−1 , Cp = 1.000 g L−1 , pH = 3, T = 25◦ C.

0.6
1.0

0.9 0.4
C/Co

0.8
2-CP
0.2
TOC
0.7
H2O2
Fe
0.6 0.0
0 30 60 90 120
Time (min)

Figure 12.13. Dimensionless concentration of reactants, products and iron vs. time for Fe-PILC6 . Catalyst
load = 1.0 g L−1 , C2-CP = 50 mg L−1 , Cp = 1.000 g L−1 , pH = 3, T = 25◦ C.

the Fe-PILCs with a really lower 2-CP removal rate in the beginning, and a subsequent acceleration
of the degradation rate. It should be noted that in this case the amount of H2 O2 required to obtain
these results was 50% higher than that employed in the case of Fe-PILCs catalysts.
In the experiment performed with 0.5 g L−1 goethite load, 41% 2-CP removal is reached in
360 minutes, while with the highest load (2.0 g L−1 ) the removal is 63%. Should be pointed
out that, after 120 minutes of reaction, the amount of 2-CP degraded is significantly lower
(<10%) than that reached with Fe-PILCs for the same reaction time. As with the Fe-PILCs,
the TOC reduction was lower than the 2-CP removal (20% and 46% for 0.5 g L−1 and 2.0 g L−1
respectively). In comparison with the Fe-PILCs, stands out the considerable consumption of
H2 O2 from the beginning of the experience, 45% and 96% for 0.5 and 2.0 g L−1 , attributed to
decomposition reaction according to the global reaction (Lin and Gurol, 1998; Ortiz de la Plata
et al., 2010a):
H2 O2 → H2 O + ½O2 (12.8)
212 A. De León et al.

1.0

0.8

0.6
C/Co

2-CP
0.4 TOC
H2O2
0.2 2-CP
TOC
H2O2
0.0
0 60 120 180 240 300 360
Time (min)

Figure 12.14. Dimensionless concentration of reactants and products vs. time for goethite.
C2-CP = 50 mg L−1 , Cp = 2.5 g L−1 , T = 25◦ C. Filled symbols: Catalyst load = 2.0 g L−1 ,
pH = 3. Open symbols: Catalyst load = 0.5 g L−1 . Adapted from Applied Catalysis B: Envi-
ronmental, 95, G.B. Ortiz de la Plata, O.M. Alfano, A.E. Cassano, Decomposition of
2-chlorophenol employing goethite as Fenton catalyst II: Reaction kinetics of the hetero-
geneous Fenton and photo-Fenton mechanisms, 14–25, Copyright (2010), with permission
from Elsevier.

This is a competing reaction with the ones leading to the generation of HO• radicals and has
been reported to occur proportionality with the catalyst loading (Ortiz de la Plata et al., 2010b;
Lin and Gurol, 1998). Should be noted that in the tests conducted at ambient temperature the
dissolved Fe in the reaction medium did not exceed 0.1 mg L−1 , which indicates the low leaching
verified with the goethite used during this experimental tests.
Figure 12.15 shows the results of 2-CP degradation with 18 mg L−1 nZVI load. The curve of
2-CP degradation presents a profile similar to the observed in the other catalysts, reaching 88%
removal at 180 minutes of reaction. At this time, the TOC removal was rather low (9%), while
H2 O2 consumption was about 18% (Ortiz de la Plata et al., 2012).
In addition to the greater efficiency in the degradation of 2-CP there is the advantage of low
peroxide consumption, being enough to work with ratios lower than those employed with goethite,
which implies a more efficient use of the oxidant by this solid maintaining the Fe dissolved levels
in really low values (below the detection limit).
A comparison between the performances of the solids shows significant differences in the
efficiency in 2-CP conversion, as well as in the oxidant use and iron stability in the solid. In
Table 12.2 can be observed that the use of nZVI allowed reaching a high 2-CP conversion, with
an amount of Fe significantly lower than the used in the other two cases. It is also observed a low
peroxide consumption, implying an efficient oxidant use by the solid, which is important from
the viewpoint of process economics. Note that TOC conversion is observed for all the solids but
in different proportions for each one. The higher removal takes place in the experience with the
higher goethite load. The requirements of hydrogen peroxide and the amount of iron present in
the experiences are really different for each catalyst. When comparing the performances of the
catalysts it is observable that nZVI has a lower H2 O2 requirement than goethite and Fe-PILCs for
reaching high 2-CP conversions. However, it can be noted that the TOC conversion was the lowest
among the employed catalysts. Beyond this better performance observed, it is necessary to bear
in mind that the accepted reaction mechanism in the case of nZVI involves the transformation
of the solid from the zerovalent state into iron oxides (Morgada et al., 2009; Ortiz de la Plata
Water decontamination by heterogeneous photo-Fenton processes 213

1.0

0.8

0.6
C/Co

2-CP
0.4
TOC

H2O2
0.2

0.0
0 60 120 180
Time (min)

Figure 12.15. Dimensionless concentration of reactants and products vs. time for nZVI. Catalyst
load = 0.18 g L−1 . C2-CP = 50 mg L−1 , Cp = 650 mg L−1 , T = 25◦ C. Reprinted from Jour-
nal of Photochemistry and Photobiology A: Chemistry 233, G.B. Ortiz de la Plata, O.M.
Alfano, A.E. Cassano, 2-Chlorophenol degradation via photo Fenton reaction employing
zerovalent iron nanoparticles, 53–59, Copyright (2012), with permission from Elsevier.

Table 12.2. Experimental conditions and results for photo-Fenton assays with different catalysts. Fe: iron
present in the suspension; C2-CP and Cp : 2-CP and H2 O2 initial concentration respectively;
X2-CP , Xp and XTOC : 2-CP, H2 O2 and TOC final conversion respectively; Time: duration of the
assay.

Catalyst Fe X 2-CP Xp X TOC Time


Catalyst load (g L−1 ) (mg L−1 ) C p /C 2-CP C p /Fe (%) (%) (%) (h)

Fe-Pilc18 1.0 176 20 5.7 99 30 15 2


Fe-Pilc6 1.0 61 20 16.4 24 9 14 2
Goethite 2.0 1257 30 1.2 63 96 46 6
Goethite 0.5 314 30 4.8 41 45 20 6
nZVI 0.2 18 13 36.1 88 18 9 3

et al., 2012), what irreversibly alters its properties. This makes a big difference with goethite and
Fe-PILCs, that promise to enable the reuse of the solid due to the maintenance almost unchanged
of its structures. Future studies will be necessary to evaluate the efficiency of the reuse of the
solids.
The evolution of the 2-CP removal differs markedly from the reported in Section 12.4.1 for
dyes degradation. Firstly, in can be noted a slow 2-CP removal in a first reaction phase which
can be associated with the low adsorption and/or degradation rate by catalytic mechanisms. The
negligible 2-CP adsorption on the Fe-PILCs was verified in tests, whose results are not included
in this study, and is reported for goethite in other works of the authors and specific literature
(Kwan and Voelker 2004; Ortiz de la Plata et al., 2010a).
In the case of the Fe-PILCs, it can be ascribed to the absence of groups of positive charge in
the 2-CP molecule, which would favor interactions with the negative residual charge of the clay
sheets. The low initial rate of the photo-Fenton reactions can be associated to the low activation
214 A. De León et al.

of the 2-CP over the active sites and/or the low amount of HO• radicals in the reaction medium,
which corresponds with the low consumption of H2 O2 observed.
The progressive increase in 2-CP removal rate can be explained by the leaching of Fe, as is
evident in Figures 12.12 and 12.13, favoring the classical mechanism of homogeneous photo-
Fenton reactions of degradation. The dissolution of Fe would be favored by the low pH and the
presence of certain reaction intermediates (Chen and Pignatello, 1997; Ortiz de la Plata et al.,
2010b; 2010c; 2012).

12.5 CONCLUSIONS

The results show the efficiency of the different iron-containing solids in heterogeneous photo-
Fenton processes. Different mechanisms involving the participation of Fe either immobilized or
in solution, participate in the degradation of the tested contaminants. It was verified that both
the structure of the solid and the model molecules tested (and their physicochemical properties)
largely determine the contribution of different mechanisms to the overall process and the influence
of pH. Reactants adsorption on the solids surface promotes the catalytic action of iron immobilized
sites, contributing to the heterogeneous mechanism. The clay aggregate size used for Fe-PILCs
preparation determines changes in their catalytic performance on dyes degradation. This could be
attributed to differences in their textural properties that determines diffusional restrictions and/or
the competition of pollutants and H2 O2 for HO• radicals.
In comparison with the goethite and nZVI solids, Fe-PILCs require lower H2 O2 consumption
to attain high 2-CP removals, although with a higher level of Fe leaching.
Additional experimental studies, which are being carried on in our groups, will provide new
evidences about the nature of the main reaction mechanism and the performance of these iron
containing solids as heterogeneous photo-Fenton catalysts.

ACKNOWLEDGEMENTS

The authors thank Comisión Sectorial de Investigación Científica, Programa de Desarrollo de


las Ciencias Básicas and Agencia Nacional de Investigación e Innovación from Uruguay, and
Universidad Nacional del Litoral, Consejo Nacional de Investigaciones Científicas y Técnicas
and Agencia Nacional de Promoción Científica y Tecnológica from Argentina for their financial
support.
The authors gratefully thank Mr. Antonio Negro for the technical assistance during Lic. De
Leon’s internships at Santa Fe. Also Eng. Susana Gervasio is acknowledged for the AA analysis
of the samples of the works made at Santa Fe and Chem. Eng. Eduardo Speranza for edition of
the English manuscript.

REFERENCES

Ameta, R., Kumar, A., Punjabi, P.B. & Ameta, S.C.: Advanced oxidation processes. In: D.G. Rao,
R. Senthikumar, J. Anthony Byrne & S. Feroz (eds): Wastewater treatment advanced processes and
technologies, CRC Press, 2012, Chap. 4, pp. 61–106.
Ayodele, O.B., Lim, J.K. & Hameed, B.H.: Pillared montmorillonite supported ferric oxalate as heterogeneous
photo-Fenton catalyst for degradation of amoxicillin. Appl. Catal. A: Gen. 413–414 (2012), pp. 301–309.
Bautista, P., Mohedano, A.F., Casas, J.A., Zazo, J.A. & Rodriguez, J.J.: An overview of the application of
Fenton oxidation to industrial wastewaters treatment. J. Chem. Technol. Biot. 83 (2008), pp. 1323–1338.
Bolton, J.R.: Ultravioleta handbook. 2nd ed., Bolton Photosciences Inc., 2001.
Bouras, O., Bollinger, J., Baudu, M. & Khalaf, H.: Adsorption of diuron and its degradation products from
aqueous solution by surfactant-modified pillared clays. Appl. Clay Sci. 37 (2007), pp. 240–250.
Water decontamination by heterogeneous photo-Fenton processes 215

Catrinescu, C., Arsene, D. & Teodosiu, C.: Catalytic wet hydrogen peroxide oxidation of para-chlorophenol
over Al/Fe pillared clays (AlFePILCs) prepared from different host clays. Appl. Catal. B: Environ. 101
(2011), pp. 451–460.
Centi, G. & Perathoner, S.: Catalysis by layered materials: A review. Micropor. Mesopor. Mat. 107 (2008),
pp. 3–15.
Chen, J. & Zhu, L.: Oxalate enhanced mechanism of hydroxyl-Fe-pillared bentonite during the degradation
of Orange II by UV-Fenton process. J. Hazard. Mat. 185 (2011), pp. 1477–1481.
Chen, Q., Wu, P., Dang, Z., Zhu, N., Li, P., Wu, J. & Wang, X.: Iron pillared vermiculite as a heterogeneous
photo-Fenton catalyst for photocatalytic degradation of azo dye reactive brilliant orange X-GN. Sep. Purif.
Technol. 71 (2010), pp. 315–323.
Chen, R. & Pignatello, J.J.: Role of quinone intermediates as electron shuttles in Fenton and photoassisted
Fenton oxidations of aromatic compounds. Environ. Sci. Technol. 31 (1997), pp. 2399–2406.
Cheng, M., Song, W., Ma, W., Chen, Ch., Zhao, J., Lin, J. & Zhu, H.: Catalytic activity of iron species in
layered clays for photodegradation of organic dyes under visible irradiation. Appl. Catal. B: Environ. 77
(2008), pp. 355–363.
De León, A., Castiglioni, J., Bussi, J. & Sergio, M.: Catalytic activity of an iron pillared montmorillonite
clay mineral in a heterogeneous photo-Fenton process. Catal. Today 133–135 (2008), pp. 600–605.
De León, M.A., Sergio, M., Bussi, J., Ortiz de la Plata G.B., Cassano, A.E. & Alfano, O.M.: Degradación de
2-clorofenol mediante el proceso foto-Fenton heterogéneo, empleando arcillas pilareadas con hierro. Con-
greso Internacional de Ciencia y Tecnología Ambiental, y I Congreso Nacional de la Sociedad Argentina
de Ciencia y Tecnología Ambiental, Mar del Plata, Argentina, 2012.
Embaid, B.P., Biomorgi, J.G., Gonzalez-Jimenez, F., Perez-Zurita, M.J. & Scott, C.E.: Using Fe-PILC as
catalyst. Appl. Catal. A: Gen. 400 (2011), pp. 166–170.
Feng, J., Hu, X., Yue, P.L., Zhu, H.Y. & Lu G.Q.: A novel laponite clay-based Fe nanocomposite and
Its photo-catalytic activity in photo-assisted degradation of Orange II. Chem. Eng. Sci., 58 (2003a),
pp. 679–685.
Feng, J., Hu, X., Yue, P.L., Zhu, H.Y. & Lu, G.Q.: Discoloration and mineralization of Reactive Red HE-3B
by heterogeneous photo-Fenton reaction. Water Res. 37 (2003b), pp. 3776–3784.
Feng, J., Hu, X. & Yue, P.L.: Discoloration and mineralization of Orange II using different heterogeneous
catalysts containing Fe: A comparative study. Environ. Sci. Technol. 38 (2004a) pp. 5773–5778.
Feng, J., Hu, X. & Yue, P.L.: A novel bentonite clay-based Fe-nanocomposite as a heterogeneous cata-
lyst for photo-Fenton discoloration and mineralization of Orange II. Environ. Sci. Technol. 38 (2004b),
pp. 269–275.
Feng, J., Hu, X., Yue, P.L. & Qiao, S.: Photo Fenton degradation of high concentration Orange II (2 mM)
using catalysts containing Fe: A comparative study. Sep. Purif. Technol. 67 (2009), pp. 213–217.
Fenton, H.J.H.: Oxidation of tartaric acid in presence of iron. J. Chem. Soc. 65 (1894), pp. 899–910.
Galeano, L.A., Vicente, M.A. & Gil, A.: Treatment of municipal leachate of landfill by Fenton-like heteroge-
neous catalytic wet peroxide oxidation using an Al/Fe-pillared montmorillonite as active catalyst. Chem.
Eng. J. 178 (2011), pp. 146–153.
Gil, A. & Gandía, L.M.: Recent advances in the synthesis and catalytic applications of pillared clays. Catal.
Rev., 42 (2000), pp. 145–212.
Hang, P.T. & Brindley, G.W.: Methylene Blue absorption by clay minerals. Determination of surface areas
and cation exchange capacities (clay-organic studies xviii). Clay. Clay Miner. 18 (1970), pp. 203–212.
Hatakeyama, T. & Liu, Z.: Handbook of thermal analysis. John Wiley & Sons. Chichester, England 1998,
pp. 251–256.
Herney-Ramirez, J., Miguel, A.V. & Madeira, L.M.: Heterogeneous photo-Fenton oxidation with pillared
clay-based catalysts for wastewater treatment: A review. Appl. Catal. B: Environ. 98 (2010), pp. 10–26.
Hou, M., Ma C., Zhang, W., Tang, X., Fan, Y. & Wan, H.: Removal of rhodamine B using iron-pillared
bentonite. J. Hazard. Mater. 186 (2011), pp. 1118–1123.
Huang, Ch.-P. & Huang, Y.-H.: Application of an active immobilized iron oxide with catalytic H2 O2 for
the mineralization of phenol in a batch photo-fluidized bed reactor. Appl. Catal. A: Gen. 357 (2009),
pp. 135–141.
Iurascu, B., Siminiceanu, I., Vione, D., Vicente M.A. & Gil, A.: Phenol degradation in water through a het-
erogeneous photo-Fenton process catalyzed by Fe treated laponite. Water Res. 43 (2009), pp. 1313–1322.
Kloprogge, J.T.: Synthesis of smectites and porous pillared clays catalysts: A review. J. Porous Mat. 5 (1998),
pp. 5–41.
Kwan, W.P. & Voelker, B.M.: Influence of electrostatics on the oxidation rates of organic compounds in
heterogeneous Fenton systems. Environ. Sci. Technol. 38 (2004), pp. 3425–3431.
216 A. De León et al.

Legrini, O., Oliveros, E. & Braun, A.M.: Photochemical process for water treatment. Chem. Rev. 93 (1993),
pp. 671–698.
Lin, S.S. & Gurol, M.D.: Catalytic decomposition of hydrogen peroxide on iron oxide: Kinetics, mechanism,
and implications. Environ. Sci. Technol. 32 (1998), pp. 1417–1423.
Liou, R.M., Chen, S.H., Hung, M.Y., Hsu, C.S. & Lai, J.Y.: Fe (III) supported on resin as effective catalyst
for the heterogeneous oxidation of phenol in aqueous solution. Chemosphere 59 (2005), pp. 117–125.
Martin-Luengo, M.A., Martins-Carvalho, H., Ladriere, J. & Grange, P.: Fe(III)-pillared montmorillonites:
Preparation and characterization. Clay Miner. 24:3 (1989), pp. 495–504.
Martínez, F., Calleja, G., Melero, J.A., Molina, R.: Heterogeneous photo-Fenton degradation of phenolic
aqueous solutions over iron-containing SBA-15 catalyst. Appl. Catal. B: Environ. 60 (2005), pp. 181–190.
Moreno, S., Sun Kou, R. & Poncelet, G.: Influence of preparation variables on the structural, textural, and
catalytic properties of Al-pillared smectites. J. Phys. Chem. B 101 (1997), pp. 1569–1578.
Morgada, M.E. Levy, I.K., Salomone, V. Farías, S.S., López, G. & Litter, M.I.: Arsenic (V) removal with
nanoparticulate zerovalent iron: Effect of UV light and humic acids. Catal. Today 143 (2009), pp. 261–268.
Najjar, W., Azabou, S. Sayadi, S. & Ghorbel, A.: Catalytic wet peroxide photo-oxidation of phenolic olive
oil mill wastewater contaminants Part I. Reactivity of tyrosol over (Al–Fe)PILC. Appl. Catal. B: Environ.
74 (2007), pp. 11–18.
Noorjahan, M., Durga Kumari, V., Subrahmanyam, M. & Panda, L.: Immobilized Fe(III)-HY: An efficient
and stable photo-Fenton catalyst. Appl. Catal. B: Environ. 57 (2005), pp. 291–298.
Noya, C., De León, A., Sergio, M. & Bussi, J.: pH influence on photo-Fenton like processes using Fe-PILCs
as catalyst. Avances en Ciencias e Ingeniería 2 (2011), pp. 35–45.
Ortiz de la Plata, G.B., Alfano, O.M. & Cassano, A.E.: Optical properties of goethite catalyst for heteroge-
neous photo-Fenton reaction. Comparison with a titanium catalyst. Chem. Eng. J. 137 (2008), pp. 396–410.
Ortiz de la Plata, G.B., Alfano, O.M. & Cassano, A.E.: The heterogeneous photo-Fenton reaction using
goethite as catalyst. Water Sci. Technol. 61:12 (2010a), pp. 3109–3116.
Ortiz de la Plata, G.B., Alfano, O.M. & Cassano, A.E.: Decomposition of 2-CP employing goethite as Fenton
catalyst. I. Proposal of a feasible combined reaction scheme of heterogeneous and homogeneous reaction.
Appl. Catal. B: Environ. 95 (2010b), pp. 1–13.
Ortiz de la Plata, G.B., Alfano, O.M. & Cassano, A.E.: Decomposition of 2-CP employing goethite as Fenton
catalyst. II. Reaction kinetics of the heterogeneous Fenton and photo-Fenton mechanisms. Appl. Catal.
B: Environ. 95 (2010c), pp. 14–25.
Ortiz de la Plata G.B., Alfano O.M. & Cassano A.E.: 2-Chlorophenol degradation via photo Fenton reaction
employing zerovalent iron nanoparticles. J. Photochem. Photobiol. A 233 (2012), pp. 53–59.
Pignatello, J.J.: Dark and photoassisted Fe3+-catalyzed degradation of chlorophenoxy herbicides by hydrogen
peroxide. Environ. Sci. Technol. 26 (1992), pp. 944–951.
Ramirez, J.H., Maldonado-Hodar, F.J., Perez-Cadenas, A.F., Moreno-Castilla, C., Costa C.A. & Madeira,
L.M.: Azo-dye Orange II degradation by heterogeneous Fenton-like reaction using carbon-Fe catalysts.
Appl. Catal. B: Environ. 75 (2007), pp. 312–323.
Rodríguez, A, Ovejero, G., Sotelo, J.L., Mestanza, M. & García, J.: Heterogeneous Fenton catalyst supports
screening for mono azo dye degradation in contaminated wastewaters. Ind. Eng. Chem. Res. 49 (2010),
pp. 498–505.
Rouquerol F., Rouquerol J. & Sing K.S.W.: Adsorption by powders & porous solids, principles, methodology
and applications. Academic Press, London, UK, 1999 pp. 110, 204, 375.
Safarzadeh-Amiri, A., Bolton, J.R. & Cater, S.R.: The use of iron in advanced oxidation processes. J. Adv.
Oxid. Technol. 1 (1996), pp. 18–26.
Son H.-S., Im J.-K. & Zoh K.-D.: A Fenton-like degradation mechanism for 1,4-dioxane using zero-valent
iron (Fe0) and UV light. Water Res. 43 (2009), pp. 1457–1463.
Sum, O.S.N., Feng, J., Hu, X. & Yue, P.L.: Photo-assisted Fenton mineralization of an azo-dye acid black
1 using a modified laponite clay-based Fe nanocomposite as a heterogeneous catalyst. Top. Catal. 33
(2005), pp. 233–242.
Yamanaka, S., Doi, T., Sako, S., Hattori, M.: High surface area solids obtained by intercalation of iron oxide
pillars in montmorillonite. Mater. Res. Bull. 19 (1984), pp. 161–168.
Wang Y., Liu C.S., Li F.B., Liu C.P. & Liang J.B.: Photodegradation of polycyclic aromatic hydrocarbon
pyrene by iron oxide in solid phase. J. Hazard. Mat. 162 (2009), pp. 716–723.
Zhang, S. Liang, S., Wang, X., Long, J., Li, Z. & Wu, L.: Trinuclear iron cluster intercalated montmorillonite
catalyst: Microstructure and photo-Fenton performance. Catal. Today 175 (2011), pp. 362–369.
CHAPTER 13

Modified montmorillonite in photo-Fenton and adsorption


processes

Lucas M. Guz, Melisa Olivelli, Rosa M. Torres Sánchez,


Gustavo Curutchet & Roberto J. Candal

13.1 INTRODUCTION

Nowadays environmental pollution is one of the major and more urgent problems to be resolved
in the world. In particular, water contamination affects most of the planet being the situation more
dramatic in the less developed countries. All the human activities from agriculture to the fabrica-
tion of sophisticated electronic equipment need water, generating enormous amounts of effluents
that should be purified before being discharge to the environment. Biological water treatment is
without doubts the most popular in the entire world because it is versatile, cheap and can be used
in big or small cities and for several industries. However, human activity generates wastewaters
containing soluble metals and/or synthetic organic compounds that are non-biodegradable or,
unlike natural occurring compounds, extremely resistant to biodegradation by native microor-
ganisms (Ali, 2010). Synthetic dyes belong to the group of poorly biodegradable compounds
and approximately 20% of the synthesized dyes are discharged in aqueous effluents without any
treatment (O Neill et al., 1999). Synthetic dyes are made up of complex aromatic molecular
structures purposely designed to resist the exposure to light, water, air, soap and oxidizing agents.
Consequently, dyes are commonly resistant to conventional biological treatment and in particular
to aerobic digestion (Asgher, 2012). Although anaerobic degradation of dyes is reported, toxic
amines intermediates are usually produced as byproducts and the coupling of anaerobic with
aerobic treatment is usually recommended (Hosseini Koupaie et al., 2013; Singh, 2011).
Physicochemical water treatments that can be used to remove non-biodegradable compounds
include adsorption, coagulation-flocculation, membrane filtration, ozonization and advanced
oxidation, between others. The sorption methods are relatively fast, can be used to remove
metals and organic compounds, and may be cheaper than others. However, they have the prob-
lem of the further elimination of the sorbed pollutant, which has to be treated as a dangerous
waste. Other techniques (like membrane filtration or ozonization) require significant amount
of electricity and the cost may be too high in regions were electricity is expensive. Advanced
oxidation processes (AOPs) are an alternative that works well for the elimination of organics.
The most used AOPs are Fenton like, photo-Fenton like and TiO2 -photocatalysis. The three pro-
cesses have been used for the degradation of dyes; Fenton like and photo-Fenton like processes
are being increasingly used in the treatment of industrial wastewater including colored waters
(Bautista et al., 2007; Lofrano et al., 2007; Meriç et al., 2005). Although the Fenton reagent
has been known for more than a century and is shown to be a powerful oxidant, the mechanism
of the Fenton reaction is still under intense and controversial discussion. However, it is gener-
ally accepted that the reaction between H2 O2 and Fe2+ in an acidic aqueous medium (pH ≤ 3)
produces HO• radicals that act as non-selective oxidants that can degrade or even mineralize
different types of organic substances. In a typical approach, four steps should be followed: pH
adjustment to 2.8–3.0, oxidation reaction, neutralization, and coagulation. The process can be
summarized as:
Fe2+ + H2 O2 → Fe3+ + HO− + HO• (13.1)
217
218 L.M. Guz et al.

As iron (II) acts as a catalyst, it has to be regenerated, which seems to occur through the
following scheme:
Fe3+ + H2 O2 ↔ Fe–OOH2+ + H+ (13.2)
Fe–OOH2+ → Fe2+ + HO•−
2 (13.3)
The photo-Fenton process produces more hydroxyl radicals in comparison to the conventional
Fenton method, thus promoting the degradation of organic pollutants. The photo-Fenton reaction
involves irradiation with UV light that significantly increases the rate of contaminant degradation
by stimulating the reduction of Fe(III) to Fe(II). Further hydroxyl radicals may be produced
via direct H2 O2 /UV photolysis (slow reaction, depending on the emission of the lamp) and the
reaction of H2 O2 with Fe2+ produced by photoreduction of Fe(III):

Fe(OH)2+ + hν → Fe2+ + HO• (13.4)

Fe2+ + H2 O2 → Fe(OH)2+ + HO• (13.5)

The photoreduction of Fe(OH)2+ takes place with wavelength of 360 nm (UVA radiation),
meaning that solar light can be used as a irradiation source in photo-Fenton process (Herney-
Ramirez et al., 2010). The higher production of HO• due to the combination of oxidants and
metallic catalysts in the presence of UVA radiation, and the potential applicability of sunlight as
UVA light source are some attractive advantages of this process (Pignatello et al., 2006). The
Fenton reagent usually incorporates Fe(II) as catalyst, but, in some cases, using only iron, it is
not possible to remove all the intermediate compounds generated as a consequence of the partial
oxidation of the pollutants (Primo et al., 2008). Other transition metals appear as an alternative to
achieve higher mineralization efficiencies; these alternatives are the so-called Fenton like process.
Copper undergoes Fenton and photo-Fenton type reactions and lead to the oxidation of several
compounds (Anipsitakis and Dionysiou, 2004; Neamtu et al., 2003; Shah et al., 2003).
As mentioned before, the photo-Fenton reaction has been widely applied under homogeneous
conditions. It is an effective method for the removal of a high variety of organic contaminants,
with the advantages of working under simple operation conditions and the relatively low price
and environmental impact of the oxidation reagents. However, this application has various serious
disadvantages, for example, the formation of Fe-containing sludge and the low pH necessary
for optimal performance. Besides, the catalytic activity of Fe(III) can be diminished by the
presence of complexing agents able to form very stable coordination compounds. The use of
supported catalysts has been reported as a solution to these problems (Herney-Ramirez et al.,
2010). Several alternatives were proposed as heterogeneous catalysts for Fenton process; bulk
catalysts containing iron may be considered, as hematite, magnetite, goethite, etc. A different
approach is the incorporation of iron into different supports as activated carbon (Dhaouadi and
Adhoum, 2010), polymers (Lee et al., 2010), zeolites (Idel-Aouad et al., 2011), clays, etc. The
use of clays as catalyst support for Fenton reaction has been recommended by several authors,
with application in general wastewater treatment (Herney-Ramirez et al., 2010) and especially
in wastewaters containing different types of dyes (Soon and Hameed, 2011). Clays have several
advantages: low cost, high surface area, low environmental impact and, in some cases, they
contain natural iron that may work as catalyst.
It was suggested that the reaction mechanism in heterogeneous catalysis involve the Fe(III) on
the surface of the support and the Fe2 O3 present on the support:

Fe3+ -on the surface of Fe_Clay + hν → Fe2+ -on the surface of Fe_Clay (13.6)

Fe2+ -on the surface of Fe_Clay + H2 O2 → Fe3+ -on the surface of Fe_Clay + HO• + OH−
(13.7)

Fe-Clay-dye + HO• → reaction intermediates → CO2 + H2 O (13.8)


Modified montmorillonite in photo-Fenton and adsorption processes 219

Fe2 O3 in Fe-Clay + dye + hν → [dye• . . . Fe2 O3] in Fe-Clay → dye•+ + Fe2 O3 in Fe-Clay + e−
CB

(13.9)

dye•+ → reaction intermediates → CO2 + H2 O (13.10)

Catalytic activity is also originated from a homogeneous process due to iron leaching from
the catalyst. Fe2+ in solution can be obtained by the photoreduction of ≡Fe3+ under UV radia-
tion. In addition, Fe3+ in solution was also produced when ≡Fe2+ is oxidized by H2 O2 . Other
transition metal cations can potentially be used as supported catalysts for Fenton or photo-Fenton
processes, but only few examples were reported at present (Timofeeva et al., 2009; Yip et al.,
2007).
Clays, and specially bentonites, deserve special attention due to their versatility for environ-
mental applications. In several countries as for example Argentina, there are rich natural sources
of bentonites that can be used in environmental applications helping the development of the
local economies. Montmorillonite is a bentonite with great specific surface area and optimal
properties for metal adsorption (Bhattacharyya and Gupta, 2008; Gu et al., 2010; Ijagbemi et
al., 2009). In particularly, the uptake of uranium and copper from aqueous solutions by the
use of montmorillonites was recently demonstrated (Bhattacharyya and Gupta, 2007; Schlegel
et al., 2009). An alternative that was recently investigated has to do with the combination of
biomass with clays. Biomass in general and microscopic biomass in particular are very active
in the adsorption of metals from solution, allowing concentrating heavy metals from diluted
solutions (Shinde et al., 2012; Yipmantin et al., 2011; Volesky, 2007). The hybrid materi-
als produced by combination of both adsorbents seem to be more efficient for the adsorption
and removal of metals in solution and to have better coagulation properties. In particular, good
results were obtained for the removal of uranium and copper in water using bacterial or fungal
biomass respectively (Ohnuki et al., 2005; Olivelli et al., 2013). When raw clays and mod-
ified clays are used to remove metals form solution, a new solid waste containing metals is
generated, which should be appropriately disposed. As an alternative, if the clays are used to
adsorb metals with catalytic activity, they can be used in oxidation processes like Fenton or
photo-Fenton.
In this work, we present a comparative study about the use of Fe(III) and Cu(II) supported
on montmorillonite as catalyst for the discoloration and mineralization of reactive Orange 16
by photo-Fenton process. Copper containing montmorillonite and bio-montmorillonite obtained
after removal of Cu(II) from water were also used to test their performance as catalyst in the
same reaction. Reactive Orange 16 was selected as pollutant target due to the requirement
of the local industry for technologies to remove dyes from wastewater before its discharge in
rivers.

13.2 EXPERIMENTAL SECTION

13.2.1 Materials
A bentonite sample (MMT) coming from the Argentine North Patagonia (Río Negro) was
used as raw material for all the synthesis described in this work. The MMT mineralogy was
evaluated by X-ray diffraction and chemical analysis in previous studies (Iborra et al., 2006;
Lombardi et al., 2003) and consists of Na-rich montmorillonite (>99%) with minor phases as
quartz and feldspars. Some physicochemical parameters of the raw MMT were determined else-
+
where: structural formula [(Si3.89 Al0.11 )(Al1.43 Fe3+
0.28 Mg0.30 )O10 (OH)2 ] M0.41 ; cationic exchange
capacity (CEC) = 174 meq/100 g clay, isoelectric point (IEP) at pH 2.7 and specific surface
area determined by N2 adsorption (BET method) SN2 = 34.0 m2 g−1 , and by water adsorption
Sw = 621 m2 g−1 (Magnoli et al., 2008).
220 L.M. Guz et al.

Reactive Orange 16 (RO16 -Sigma) was used as received. Hydrogen peroxide (30%), sulfuric
acid and sodium sulfite, all PA quality, from Merck, were also used. Deionized water was obtained
with an Osmoion water purifier. The formula of RO16 is given by:

13.2.2 Iron (III) modified montmorillonite (Fe-MMT)


Iron chloride hexahydrate (15 g, Aldrich) and montmorillonite (15 g) were respectively dissolved
and suspended in acetone (50 mL, Merck). The slurry was stirred for 1 hour, washed with acetone
and ethanol, and finally dried at 60◦ C overnight. The obtained product was named Fe-MMT.
To determine the iron loading, a sample of Fe-MMT was digested in a 5.0 M HCl+HNO3
acid solution at room temperature. The content of iron in the solution was determined by the
thiocyanate method. The iron load was 24 ± 1 mg/g of clay.

13.2.3 Copper (II) modified montmorillonite (Cu-MMT)


MMT fired at 500◦ C for 8 h (3.0 g) was suspended in 100 mL of a 0.050 M CuSO4 (Merck)
solution at pH 5.0, and stirred for 1 hour at 50◦ C. The solid was decanted and washed 3 times
with deionized water and dried at 60◦ C overnight. The obtained material was named Cu-MMT. To
determine the copper loading, a sample of Cu-MMT was digested in 1.0 M HNO3 acid. Cu(II) was
determined spectrophotometrically by the dithizone method. The copper loading was 60 ± 2 mg/g
of clay.

13.2.4 Biomodified montmorillonite (Apha-BMMT)


A copper and uranium resistant acidophilic fungi genus was used, Aphanocladium sp. (Apha sp.).
To generate the clay biopolymer (named as Apha-BMMT ), the biomass was grown axenically
in batch systems with P5 culture medium (1.28 g L−1 K2 HPO4 ; 3 g L−1 (NH4 )2 SO4 ; 0.25 g L−1
MgSO4 .7H2 O; 10 g L−1 glucose; 0.1 g L−1 thiamine; 1% (v/v) microelements solution) containing
MMT clay 1% (w/v) and 5% V/V of initial inoculum. Generated Apha-BMMT were recovered
by centrifugation (20 minutes, 2200 g, 4◦ C) and washed with distilled water.

13.2.5 Adsorption of Cu(II) on MMT and Apha-BMMT


MMT samples were stabilized by suspension overnight in P5 culture medium, followed by washing
with deionized water, centrifugation and drying at 60◦ C. This material, named P5-MMT, was used
as sorbent for Cu(II) dissolved in water solutions.
Copper sulfate solutions containing from 0.16 to 7.9 mM copper pH 3.6 were prepared. Dupli-
cate Apha-BMMT or P5-MMT samples (0.1 g) were shaken during 2.25 h (according to previous
equilibrium essays, data not shown) in polypropylene tubes with 10 mL aliquots of each solution.
After the equilibration time, two samples of 1.0 mL were taken from each tube and centrifuged
(5000 g, 8 min). The supernatant was analyzed for copper with the Cuprizone technique, and the
solid was reserved for XRD and electrophoretic mobility analysis. The metal adsorbed was calcu-
lated as the difference between the initial concentration and that of the supernatant in equilibrium.
Modified montmorillonite in photo-Fenton and adsorption processes 221

The copper adsorption capacity was fitted by Langmuir, Freundlich and sigmoidal isotherms using
the SigmaPlot 10.0 software (Statistics – Regression Wizard).
The same procedure was followed to obtain P5-MMT and Apha-BMMT saturated with Cu(II)
for testing their catalytic activity in photo-Fenton experiments for the degradation of RO16. The
saturated powders were washed several times by centrifugation (until Cu(II) was not detected
in the rinsed water) and dried at 60◦ C. The obtained catalysts were named P5-Cu and Apha-Cu
respectively.

13.2.6 Materials characterization


Samples of the different materials were characterized by X-ray diffraction (XRD, Siemens
D-5000) and by measurements of electrophoretic mobility in KCl 1.0 mM at different pH (EPM,
Brookhaven 90-Plus).

13.2.7 Photo-Fenton experiments


Photo-Fenton experiments were performed in a batch recirculating system (1.0 L min−1 flow rate)
consisting of an annular glass reactor (415-mm length, 32-mm internal diameter), a peristaltic
pump (APEMA BS6, 50 W), and a thermostatted (25◦ C) cylindrical reservoir. A blacklight tubular
UV lamp (Philips TLD/08, 15 W, 350 nm < λ < 410 nm, maximum emission at 366 nm) was
installed inside the annular reactor as the illumination source. Scheme 13.1 shows a flow diagram
of the system. The total volume of the circulating mixture was 450 mL, of which 100 mL was inside
the photoreactor and constantly irradiated. The catalyst (1 g L−1 ) was dispersed by continuous
stirring in 450 mL of a water solution containing 100 mg L−1 of RO16 and the pH adjusted to 3.0
with diluted sulfuric acid. It was experimentally determined that adsorption of RO16 on MMT, Fe-
MMT, Cu-MMT, P5-Cu or Apha-Cu was negligible. After 10 min of stabilization under stirring,
the pump and the light were turned on, letting the suspension circulate for the reactor for at least
10 min. A 10 mL sample was taken, stabilized with 2 mL of 1.0 M sodium sulfite and stored at
4.0◦ C. Hydrogen peroxide, 30%, was added to the system in order to obtain the desired H2 O2
initial concentration. During irradiation, the suspension was vigorously stirred in the reservoir.
Samples were taken at certain intervals to determine RO16 and TOC concentrations. The samples
were immediately quenched by addition of sodium sulfite in the proportion given above. The solid
catalyst was removed from the suspension by centrifuging at 10000 rpm in 2 mL plastic tubes.
RO16 concentration was determined by UV-Vis spectrophotometry with an Ocean optics
HR2000 spectrophotometer. Total organic carbon (TOC) was measured with a Shimadzu 5000-A
TOC analyzer in the nonpurgeable organic carbon (NPOC) mode. In what follows, we refer to
this measurement simply as TOC. Actinometric measurements were performed with ferrioxalate.
A photon flow per unit volume (q0n,p /V, where q0n,p is the incident photon flux and V is the irradi-
ated volume) of 7.4 µ Einstein s−1 L−1 was calculated for the black-light lamp (assuming 366-nm
monochromatic light).

13.3 RESULTS

13.3.1 Adsorption of Cu(II) on P5-MMT and Apha-BMMT


Figure 13.1 shows copper adsorption isotherms for P5-MMT and Apha-BMMT.
At low copper concentrations (lower than 2 mM), all the samples show similar adsorption
capacity. From 2 mM, the adsorption capacity rises, mainly in the systems with biomass. This
sigmoidal shape of isotherms suggested different adsorption sites with different affinities for
copper or the existence of a cooperation effect that increases the affinity between copper (II) and
adsorption sites. Although the biopolymer had lower adsorption capacity of the biomass or MMT,
the coagulation capacity of the system was increased, leading to an improved separation of solids
from the solution.
222 L.M. Guz et al.

Scheme 13.1. Photo-reactor system flow diagram.

Figure 13.1. Adsorption isotherms of Cu(II) on different substrates.

13.3.2 Catalysts characterization


Figure 13.2A shows XRD patterns of the three different solids: MMT, Fe-MMT and Cu-MMT.
The d001 reflection is the main source to identify the clays (interlayer space). Pure MMT display
a diffraction peak at 6.6◦ that corresponds to a d001 basal spacing of 13.6 Å. On the other hand,
Fe-MMT shows a diffraction peak at 5.6◦ that corresponds to a d001 basal spacing of 16.5 Å. In
the case of Cu-MMT, the reflection peak shifted to 7.6◦ , which is similar than in the case of MMT
fired at 500◦ C (not shown). The d001 basal spacing in these cases is 11.6 Å that indicates the
collapse of the structure as consequence of the thermal treatment. The diminution in the intensity
of the reflection peak is also due to the disruption of the structure.
The XRD patterns of MMT and BMMT with adsorbed copper was different. A shift of the
reflexion peak d001 towards smaller values of 2θ was observed for P5-Cu with respect to that
found for P5-MMT. The shift of the reflexion peak d001 , indicated the entrance of the copper
Modified montmorillonite in photo-Fenton and adsorption processes 223

Figure 13.2. XRD Patterns of MMT, Fe-MMT and Cu-MMT (Figure 2A) and P5-MMT, P5-Cu, Apha-MMT
and Apha-Cu (Fig. 13.2B).

Figure 13.3. Electrophoretic mobility of MMT, Fe-MMT and Cu-MMT at different pH. 0.050 g L−1 of the
powders were dispersed in 1 mM KCl. pH was adjusted with HNO3 or KOH as needed.

cation to the clay interlayer. The asymmetrical shape of the reflexion peak d001 for all samples
with copper indicated the existence of a heterogeneous interlayer. To achieve some precision of
the latter behavior, the mathematical decomposition of the reflexion peak d001 for P5-Cu and
Apha-Cu was performed. Two peaks were found for the d001 reflexion peak decomposition of
both samples, which confirmed the existence of a heterogeneous interlayer originated by water
and copper cations in the interlayer. In Apha-Cu, the input of copper to the interlayer is of lesser
extent than in the case of MMT. This decrease may be due to the high affinity of the biomass by
Cu(II) that competes with the binding sites into the clay interlayer
Figure 13.3 shows the effect of pH on the electrophoretic mobility of MMT, Fe-MMT and
Cu-MMT. MMT electrophoretic mobility is almost independent of pH due to its intrinsic nega-
tive charge, which is consequence of its crystalline structure. Proton adsorption on MMT takes
place at low pH values, leading to a diminution in the absolute value of ζ-potential. In the case
of Fe-MMT, the ζ-potential notably increases as the pH decreases, while Cu-MMT displays a
behavior between MMT and Fe-MMT. The effect of Fe(III) on the ζ-potential indicates that iron
is present on the surface of the Fe-MMT particles, shifting the isoelectric point to values closer
to that corresponding to iron oxides (pH: 7.8–8.5). In the case of Cu(II), there is only a shift of
ζ-potential to less negative values, meaning that Cu(II) compensate in part the negative surface
charge of MMT.
224 L.M. Guz et al.

Figure 13.4. Temporal evolution of RO16 in water solutions. RO16 concentration: 100 mg L−1 ; MMT,
Fe-MMT concentration: 1.0 g L−1 ; pH = 3.0; [H2 O2 ]0 = 100 mM.

Table 13.1. Pseudo first order constant rate and correlation coefficients
corresponding to the experiments shown in Figure 13.3.

System k1 (s−1 ) r2

MMT + UVA + 0.1 M H2 O2 pH 3 0.36 ± 0.01 0.9891


Fe-MMT + 0.1M H2 O2 pH 3 9.2 ± 0.3 0.9968
Fe-MMT + UVA + 0.1 M H2 O2 pH 3 13.7 ± 0.2 0.9998
Fe-MMT + 0.05 M H2 O2 pH 3 10.3 ± 0.7 0.9936
Fe-MMT + UVA + 0.05 M H2 O2 pH 3 18.2 ± 0.6 0.9991

13.3.3 Photo-Fenton experiments


Figure 13.4 shows the temporal evolution of the concentration of RO16 in water solutions exposed
to different treatments that use Fe-MMT as catalyst, together with different controls.
Degradation of the dye in the presence of MMT + UV, MMT + H2 O2 or Fe-MMT+UV was
negligible. However, when Fe-MMT was used in the presence of H2 O2 or H2 O2 + UVA, the
degradation of the dye was very fast due to the occurrence of Fenton and photo-Fenton processes,
respectively. It is noticeably that photo-Fenton is slightly more efficient than Fenton. The dye was
also discolored in the presence of H2 O2 and UVA with or without MMT (only the data obtained
in the presence of MMT are shown in Fig. 13.4), but with a much lower degradation rate.
The degradation rate could be described by a first order kinetics; the pseudo first order constants
are given in Table 13.1. It is notable the increment in k1 when the Fenton experiments were run
under UVA irradiation.
Table 13.1 also shows the effect of H2 O2 concentration on the pseudo first order constant.
When the concentration decreased from 0.100 M to 0.05 M, the constant rates slightly increased.
Figure 13.5 shows the temporal evolution of RO16 water solutions exposed to different oxida-
tion treatments using Cu-MMT. The degradation rate was noticeably lower than in the experiments
run with Fe-MMT. The Fenton-like reaction with Cu-MMT was even slower than with pure
MMT. However, under UVA illumination, the degradation rate increased and the solutions were
completely bleached in 5 hours.
The color of the solution during degradation shifted to red when Cu-MMT was used as catalyst.
This effect was observed with naked eye and suggests that the mechanism of degradation is
different depending if Fe-MMT or Cu-MMt is used as catalyst in photo-Fenton process. Figure
Modified montmorillonite in photo-Fenton and adsorption processes 225

Figure 13.5. Temporal evolution of RO16 in water solutions. RO16 concentration: 100 mg L−1 ; MMT,
Cu-MMT concentration: 1.0 g L−1 , pH = 3.0; [H2 O2 ]0 = 100 mM.

Figure 13.6. UV-vis spectra of samples taken during photo-Fenton degradation 100 mg L−1 RO16, pH
3.0, catalyst concentration 1.0 g L−1 , [H2 O2 ] = 100 mM. (A) MMT, (B) Cu-MMT. Part of the
copper was released to the solution at pH 3.0; its concentration after 24 h of irradiation was
18 ± 2 mg L−1 .

13.6 shows UV-vis spectra of solutions taken at different times during the degradation by photo-
Fenton of RO16, using MMT (Fig. 13.6A) or Cu-MMT (Fig. 13.6B). In the first case, the intensity
of the band centered at 490 nm decreased with treatment time without changes in the maximum
position. In the second case, the position of the maximum shifted to higher wavelength (red) as
the intensity of the band decreased.
The diminution of the total organic carbon content (TOC) during the treatment is another
important consequence of the oxidation treatment. Figure 13.7 shows the temporal evolution of
TOC during a wide period of time. Fenton and photo-Fenton with Fe-MMT quickly reduced
TOC until approximately 65 or 57% of TOC remained in solution, respectively. Photo-Fenton
process was faster than Fenton and led to lower TOC at the end of the process. On the other hand,
Cu-MMT worked slowly at the beginning but led to lower TOC at the end of the process. These
results also show that the mechanism involved in the degradation of the dye in the presence of
Fe-MMT or Cu-MMT was different. Pure MMT also shows activity in the reduction of TOC in
solution, but the reduction rate was lower and a higher percentage of TOC remained at the end of
the experiment.
Degradation of RO16 was also performed in the presence of other copper containing catalysts
as P5Cu (Cu(II) adsorbed on MMT) and Apha Cu (Cu(II) adsorbed on BMMT), to compare
the activity of these materials (obtained as “waste” in the adsorption of Cu(II) from aqueous
solutions) with Cu-MMT, especially prepared as photo-Fenton type catalyst. Figures 13.8A and
B show discoloration rate and TOC evolution, respectively.
226 L.M. Guz et al.

Figure 13.7. TOC evolution during Fenton and photo-Fenton process using different catalysts. In all the
experiments, [H2 O2 ]0 = 0.100 M, catalyst concentration: 1.0 g L−1 , RO16 concentration:
100 mg L−1 , pH = 3.0.

Figure 13.8. Evolution of color (A) and TOC (B) during photo-Fenton process using different catalyst con-
taining Cu(II). In all the experiments [H2 O2 ]0 = 0.100 M, catalyst concentration: 1.0 g L−1 ,
RO16 concentration: 100 mg L−1 , pH = 3.0.

All the Cu(II)-containing catalysts displayed similar behavior with respect to their activity in
discoloration. The kinetics could be described by a pseudo-first order law (the parameters are
given in Table 13.2). An experiment was done as a control for homogeneous catalysis using CuSO4
as a source of Cu(II) (see Fig. 13.8A). The concentration of Cu(II) in solution (20.0 mg L−1 ) was
close to that obtained by leaching from Cu-MMT after 24 h of reaction at pH 3.0 (18 mg L−1 ).
The behavior of this control experiment was similar to those in the experiments with supported
Cu(II), indicating that most of the catalytic activity may be due to dissolved Cu(II).
Figure 13.8B shows that the evolution of TOC was similar for the systems with Cu-MMT and
P5-Cu, the system with P5-Cu being slightly faster. At the end of the experiments, the content
of TOC was similar in both cases and lower than that corresponding to pure MMT. In the period
15–24 h, TOC did not decrease in spite that there was H2 O2 in solution (after 14 h of treatment,
80% of the initial H2 O2 remained in all copper systems, measured by the VO3− method (Pupo
Nogueira et al., 2005). The situation was completely different for Apha-Cu, where TOC remained
constant during the first 5 h of treatment, but rapidly increased at longer times.
Modified montmorillonite in photo-Fenton and adsorption processes 227

Table 13.2. Pseudo first order constant rate and correlation coefficients
corresponding to the experiments shown in Figs. 13.5 and 13.8A.

System k1 (s−1 ) r2

MMT + UVA + 0.1 M H2 O2 pH 3 0.36 ± 0.01 0.9891


Cu-MMT + 0.1 M H2 O2 pH 3 0.19 ± 0.01 0.9886
Cu-MMT + UVA + 0.1 M H2 O2 pH 3 0.53 ± 0.02 0.9895
20 mg L−1 CuSO4 + UVA + 0.1 M H2 O2 , pH 3 0.59 ± 0.04 0.9829
P5-Cu + UVA + 0.1 M H2 O2 pH 3 0.62 ± 0.06 0.9638
Apha Cu + UVA + 0.1 M H2 O2 pH 3 0.60 ± 0.04 0.9771

Figure 13.9. Temporal evolution of RO16 during photo-Fenton like treatment at pH 3.0 or 6.0. RO16
concentration: 100 mg L−1 ; catalyst concentration: 1.0 g L−1 , [H2 O2 ]0 = 100 mM.

It is well established that pH has a notable effect on the catalytic activity in Fenton and photo-
Fenton processes. Figure 13.9 shows the evolution of RO16 concentration using both catalysts
at two different initial pH values: 3.0 or 6.01 . In this Figure, the difference in degradation rate
when Fe-MMT or Cu-MMT was used as catalyst is remarked; as shown in Tables 13.1 and 13.2,
all systems except Fe-MMT at pH 6.0 can be modeled by a pseudo first order kinetics, being the
constant rate more than one order higher for the Fe-MMT systems than for the Cu-MMT systems.
The effect of pH on the discoloration rate when Fe-MMT was used as catalyst was notable. At
pH 6.0, there was an induction period that delays the discoloration of the solution, but after
the first 40 minutes, discoloration occurred relatively fast. In the case of Cu-MMT, at pH 6.0
the discoloration was faster at the beginning of the treatment than at pH 3.0. As the treatment
went further, the reaction slowed down and, at the end of the treatment, both systems reached
approximately the same degree of discoloration. It is very interesting that when pure MMT was
used as catalyst, the discoloration rate was quite similar at both pH values. It should be mentioned
that in the experiments made at initial pH 6.0, the pH decreased during the experiment, being 4.5
the final value in all cases.
Figure 13.10 shows the evolution of TOC in the experiments performed at different pH values.
In the case of Fe-MMT, TOC diminution followed a pattern similar to discoloration at each pH.
At pH 6.0, TOC diminutions started after an induction period but, at the end of the experiment,

1The absorption spectrum of RO16 is independent of pH in the range 2–8.


228 L.M. Guz et al.

Figure 13.10. TOC evolution during photo-Fenton like treatment at pH 3.0 or 6.0. RO16 concentration:
100 mg L−1 ; catalyst concentration: 1.0 g L−1 , [H2 O2 ]0 = 100 mM.

TOC values were similar for both pH values. In the case of pure MMT and Cu-MMT, similar
behavior was observed at pH 3.0 and 6.0. The more remarkable result of this experiment is that
for the different catalysts the final TOC value was similar at both pH values.

13.4 DISCUSSION

The incorporation of metal cations into MMT can be done by different ways. Cation exchange of
the exchangeable cations of the clay may be the simplest way, although the content of metal cation
could be too low. Impregnation and pillaring appears as more effective methods. Also, the metals
can be simply adsorbed at the surface. Incorporation of metals into the inter-lamellar spaces of
the clays produced an increment in the d001 basal spacing that could be detected by XRD analysis.
Figure 13.1 shows clearly that Fe-MMT presented a d001 spacing larger than pure MMT (16.5
and 13.6 Å, respectively), which indicates that Fe(III) was incorporated into the MMT structure.
On the other hand, Cu-MMT was prepared by impregnation of calcined MMT. In this case, the
XRD analysis indicated that the basal spacing was smaller than in raw MMT (11.6 Å), meaning
the partial collapse of the structure (which is supported by the diminution in peak intensity).
Consequently, Cu(II) was not introduced between layers and was mostly presented at the surface
of the particles. In the case of P5-Cu, the d001 basal spacing increased indicating that some Cu(II)
was exchanged into MMT. It should be noted that the radii of Cu2+ and Cu(H2 O)2+ 6 (octahedral)
are 0.94 Å and 1.96 Å respectively (Persson, 2010), so there is plenty of room for Cu(II) in the
interlayer space of MMT. In all the cases, the metals were also adsorbed at the surface as can be
deduced from the changes in electrophoretic mobility shown in Figure 13.2. The adsorption of
metallic cations decreased the surface charge of MMT, decreasing the mobility and shifting the
isoelectric point to lower pH values. Metals at MMT surface showed catalytic activity in both
Fenton and photo-Fenton reactions. The sigmoidal shape displayed by the adsorption isotherm of
Cu(II) on BMMT could be due to the heterogeneity of sites available on this adsorbent. Besides the
interlayer spaces and the surface sites typically present in clays, BMMT also display functional
groups present in the biomass attached to the MMT. These groups provide a larger quantity
of metal binding sites and greater affinity to the system. Evidence of adsorption of copper in
interlayer space is shown by XRD analysis (Fig. 13.2B). Studies of Cu(II) interactions with fungi
and bacteria showed that Cu(II) may be associated with the functional groups on the cellular
Modified montmorillonite in photo-Fenton and adsorption processes 229

surface. FT-IR analysis indicates that copper is bound to sites in the interlayer of MMT and
amines, amides and carboxylates of proteins and carbohydrates of biomass (data not shown).
RO16 was quickly degraded by both Fenton and photo-Fenton process catalyzed by Fe-MMT.
The dye was also partially degraded by the simultaneous presence of MMT, H2 O2 and UVA; this
effect may be a consequence of sensitization of the dye by UVA absorption (see Fig. 13.6A), or
due to the leaching of a small amount of iron from the MMT that could trigger some photo-Fenton
activity. It should be noted that natural MMT has iron in its composition that can be leached to
the solution at pH 3 (0.03 mM measured by atomic absorption spectroscopy). Photo-Fenton was
a little more active than Fenton, as can be observed by comparing the values of the pseudo-first
order constants shown in Table 13.1. The degradation rate of RO16 by photo-Fenton increased
when the concentration of H2 O2 was reduced to 0.050 M. This effect may be a consequence of
the well-known scavenger effect of H2 O2 on HO• radicals represented by reactions (13.11) and
(13.12) (Pupo Nogueira et al., 2007):
HO• + H2 O2 → HO•2 + H2 O (13.11)

Hydroperoxyl radicals display a much lower redox potential (1.42 V vs. NHE) than HO• radicals
(2.73 V vs. NHE), and can be eliminated by Fe3+ in solution:
Fe3+ + HO2 → Fe2+ + O2 + H+ (13.12)

The excess of H2 O2 very likely reduced the concentration of HO• available for RO16 oxidation,
diminishing the degradation rate as the H2 O2 concentration increased. The stoichiometric ratio
dye:H2 O2 for oxidation of RO16 is 1:40. In this study, the ratio was 10 and 20 times higher in favor
to H2 O2 to allow a comparison between the experiments performed with the catalysts containing
Cu(II) or Fe(III).
When Cu-MMT was used as catalyst, discoloration of RO16 was observed in the dark and
under illumination, but the degradation rate was notably lower than in the case of Fe-MMT. In
the dark, the discoloration was even slower than with pure MMT. UVA illumination produced a
marked increment in discoloration rate; the difference in the degradation rate observed in the dark
with respect to UVA illumination was higher than in the case of Fe-MMT (compare Figs. 13.3
and 13.4). The activity of Cu(II) in Fenton and photo-Fenton processes was documented before,
although the details are less known than in the case of Fe(II)/Fe(III) (Ciesla et al., 2004; Ghiselli
et al., 2004; Masarwa et al., 1988; Sykora, 1997). Cu(II) has to be reduced to Cu(I) to enter the
typical Fenton cycle that produce HO• and regenerates the catalyst:
Cu2+ + 1 e− → Cu+ (13.13)

Cu+ + H2 O2 → Cu2+ + OH− + HO• (13.14)


The reduction of Cu(II) to Cu(I) may be mediated by reaction with organic compounds or by
reduction with HO− •−
2 or O2 (Ciesla et al., 2004; Sykora, 1997):

Cu2+ + HO− + −
2 → Cu + O2 + H
+
(13.15)

Cu2+ + O− +
2 → Cu + O2 (13.16)
It was proposed before that because copper often disproportionate H2 O2 intermediates, con-
verting them to H2 O, aqueous Cu(II) may reduce the oxidative capacity of H2 O2 (Ghiselli et al.,
2004). This effect may explain the low activity of Cu-MMT in Fenton reaction (even lower than
that of MMT + H2 O2 + UVA).
It was determined that copper can be involved in photoredox cycles that can explain their activity
in photo-Fenton processes. Copper complexes can be excited by sunlight when the ionization
energy of the ligands coordinated to Cu(II) is not too high. As consequence of the reactive decay
of the excited state by inner sphere electron transfer, the Cu(II) central atom is reduced to Cu(I),
whereas a ligand is oxidized to its radical and leaves the coordination sphere (Ciesla et al., 2004):

[CuII Lx ]2+ −−−→ [CuI Lx−1 ]+ + L+ (13.17)
230 L.M. Guz et al.

Photoproduced Cu(I) can be oxidized to Cu(II) by O2 or by H2 O2 :

Cu+ + O2 → Cu2+ + O•−


2 (13.18)

Cu+ + H2 O2 → Cu2+ + HO• + OH− (13.19)


The oxidized ligands can be involved in redox process that lead to the oxidation of the same
ligands or other organics present in the solution.
As shown in Figure 13.5, the discoloration process of RO16 in the presence of UVA light was
preceded by the formation of a red intermediate. That means that the mechanism of degradation
involves a first step of oxidation but preserving the aromatic and conjugated structure. The
reaction may involve the formation of a coordination complex between Cu(II) and the dye, which
is further photooxidized, starting the redox cycle. As shown in Figure 13.5, the dye and its colored
byproducts have absorption bands in the UVA range, which can explain its photoactivity. Studies
involving HO• scavengers to identify the effect of hydroxyl radicals in the oxidation process and
isolation of the red byproduct for chemical characterization are in course to check the previous
hypothesis. At this point, it is not clear if at pH 3.0 the Cu(II) involved in the process is the aqueous
copper or the immobilized copper. As it was explained in the Results section, similar results were
obtained when Cu-MMT was replaced by Cu(II) in a concentration equivalent to that obtained
after Cu-MMT leaching. However, as will be discussed later, at pH 6.0, Cu(II) was not leached
to the solution but photo-Fenton activity was also observed.
In the literature, there are few reports about supported copper as Fenton or photo-Fenton
catalyst. Recently, Yip et al. (2007) reported the photocatalytic activity of clay supported copper in
photo-Fenton processes, but the catalyst was prepared by chemical vapor deposition and reduced
to Cu(I); also the light source was UVC. In spite of the differences in the preparation of the
catalyst, there were several similarities with the results reported here; for example, TOC could
be successfully reduced during the treatment. By comparing the results of both works, it seems
that reduction of Cu(II) to Cu(I) produces a more active catalyst for photo-Fenton process, but a
reduction step is necessary and UVC is needed as light source.
Discoloration itself is an important result because it reduces the visual impact of the pollutants
and diminish the absorption of the incident light, which is necessary for the growing of biota
that may contribute to further elimination of byproducts. However, the total organic content is
also an important issue because it is related to the presence of recalcitrant pollutants that may
be more toxic than the dye itself (as was discussed in the introduction). The spectra shown in
Figure 13.6 indicate the elimination of aromatic rings during the degradation of RO16, as can be
deduced from the diminution of the intensity of the absorption bands centered at 385 and 290
nm. This result explains the fast discoloration of the solution exposed to the Fenton treatment. On
the other hand, the analysis of the evolution of TOC content during the degradation experiments
showed that at the beginning of the reaction TOC decreased with a similar trend than discoloration,
but mineralization stopped after TOC reached 40–60% of its initial value. Similar results were
recently reported in the degradation of other naphthalene azo dyes by the Fenton reaction (Zhu
et al., 2012). Degradation of azo dyes usually involves the elimination of the azo bound as N2 ,
followed by different steps that include OH incorporation to the aromatic ring, dealkylation and
ring opening. As it is documented in the literature, the hydroxylation of benzene and naphthalene
rings leads to ring opening, giving short-chain carboxylic acids (Stylidi et al., 2003). The low
TOC degradation is usually assigned to the generation of alkyl carboxylic acids with relatively
low molecular weight (Klamerth et al., 2009; Zhu et al., 2012). Such compounds, as acetic,
oxalic or maleic acid, are recalcitrant with respect to Fenton or photo-Fenton treatments. In the
presence of UVA, light TOC elimination was faster, as can be seen in Figure 13.7. This effect
is related with the generation of HO• by photoreaction of Fe(OH)+ 2 . As the reaction advanced,
recalcitrant byproducts formed and TOC remained stable, but at lower percentage than in the dark
experiments.
Modified montmorillonite in photo-Fenton and adsorption processes 231

In the case of the copper containing catalysts Cu-MMT and P5-Cu, the initial rate of TOC
diminution in the photo-Fenton reaction was notably lower than the corresponding to Fe-MMT.
However, TOC decreased continuously during the first 15 h of treatment, reaching at the end of
the experiment lower TOC than in the Fe-MMT case. P5-Cu resulted more active than Cu-MMT
in the elimination of TOC, probably due to the higher area of this material that was not exposed
to thermal treatment. The reduction of TOC and the discoloration of RO16 observed with P5-Cu
indicate that this material, obtained by adsorption of copper in water, can be used as catalyst for
the elimination of organic pollutants in water. Some improvements are needed in order to reduce
the secondary contamination produced by the release of copper to the solution at pH 3.0. In the
case of Apha-Cu, TOC increased after the first five hours of treatment due to the elimination of
the original biomass attached to MMT.
pH had an important role in the degradation of RO16 by photo-Fenton process under the condition
studied in this work, as it is shown in Figure 13.9. Because the studied pH values had no effect
on the degradation of the dye in the absence of catalyst (similar discoloration performance was
observed in experiments with MMT, H2 O2 and UVA), the changes in the discoloration rate at
pH 3.0 or 6.0 can be related with changes in the activity of the catalysts or in the reaction
mechanisms. It is well known that when iron species are used as catalyst, pH has an important
effect on the efficiency of homogeneous Fenton and photo-Fenton processes. The optimal pH is
close to 3. Lower pH produces less active iron species in solution (as Fe(H2 O)3+ 6 ) and decreases
the production of HO• from H2 O2 (the specie H3 O+ 2 is more reluctant to the production of
radicals); in addition, HO• are scavenged by H+ (Herney-Ramirez et al., 2010). Higher pH leads
to precipitation of FeOOH, removing iron from solution, with a consequent diminution in the
degradation rate. In iron based heterogeneous Fenton and photo-Fenton, the effect of pH follows
the same trend as in the homogeneous process, but the effects are not as dramatic due to the
immobilization of the active species. As shown in Figure 13.9, RO16 was discolored at both
studied pH values. Because iron is supported, the change in speciation seems to be less affected
by pH. The immobilized Fe(III) species cannot be easily transformed in Fe(H2 O)3+ 6 at low pH or
Fe(OH)3 at high pH. Also, due to the interaction of Fe(III) with the clay sheets, the speciation is
different and less sensitive to pH changes (De Leon et al., 2008; Najjar et al., 2007). Figure 13.9
also shows that at pH 6.0 Fe-MMT present an induction time before discoloration take place. At
present, the origins of this phenomenon are unknown but may be related with the decrease in
the pH that took place during the performance of the process (the final pH was 4.5), as reported
by Zhang et al. (2011). Lower pH leads to the release of small amounts of iron to the solution,
which may enhance the degradation rate. The time necessary to reduce the pH and leach iron may
explain the induction time.
Although discoloration rate and TOC removal started with a lower rate at pH 6.0 with respect
to 3.0, the dye was removed with an acceptable rate and the percentage of mineralization was
the same at both pH values (see Fig. 13.10). Because working at pH 6.0 is safer, demand less
preparation of the wastewater (which in the case of textiles is usually at pH over 7) and no special
acid resistant equipment is needed, it is worth to work at pH 6 in spite of some reduction in
degradation rate.
In the case of Cu-MMT, similar discoloration rate and TOC diminution rate were observed at
both studied pH values, but a small increment in the initial discoloration rate was observed at pH
6.0 and Cu(II) was not leached to the solution. Although there are relatively few antecedents in
the use of Cu(II) as catalyst, the stability of its performance with pH was reported before (Yip
et al., 2007). The reasons of this behavior are not clear but may be related with the hydrolytic
performance of Cu(II) in water. These results are very promissory for the future application
of copper based catalyst in photo-Fenton like processes, because the degree of mineralization
is usually higher than that for the iron based catalyst, they are less pH dependent and, at pH
6.0, Cu(II) leaching is very low. Besides, the adsorbents used for copper removal, and other
copper enriched wastes (Huanosta-Gutierrez et al., 2012) can be used as catalysts in Fenton like
processes.
232 L.M. Guz et al.

13.5 CONCLUSIONS

Montmorillonite (MMT) and their iron and copper modifications result attractive materials for
water decontamination. The adsorptive capacity of MMT is appropriate for removal of heavy
metals from water. Fe-MMT is very active as catalyst for the discoloration of dyes in water by
Fenton or photo-Fenton like process at initial pH 3.0 and 6.0. TOC diminution reaches approx-
imately 40% of the initial value due to the production of recalcitrant intermediates. Cu-MMT
is also active as catalyst at initial pH 3.0 and 6.0, but only in photo-Fenton like processes. The
discoloration rate is lower than that corresponding to Fe-MMT, but the degree of mineralization is
higher, reaching values close to 60% of the initial TOC. The mechanism involved in the oxidation
is different than the corresponding to Fe-MMT, leading to the production of other dyes before
discoloration. Leaching of Cu(II) may be a problem at pH 3.0 but not at pH 6.0, while the activity
is quite similar. The waste material obtained after removing Cu(II) from water was successfully
used as catalyst in photo-Fenton like processes.

ACKNOWLEDGEMENTS

This work was performed as part of FONARSEC program, FSNano 2010.

REFERENCES

Anipsitakis, G. & Dionysiou, D.: Transition metal/UV-based advanced oxidation technologies for water
decontamination. Appl. Catal. B 54 (2004), pp. 155–163.
Asgher, M.: Biosorption of reactive dyes: A review. Water Air Soil Pollut. 223 (2012), pp. 2417–2435.
Bautista, P.; Mohedano, A.F.; Gilarranz, M.A.; Casas, J.A. & Rodriguez, J.J.: Application of Fenton oxidation
to cosmetic wastewaters treatment. J. Hazard. Mater. 143 (2007), pp. 128–134.
Bayo, J.: Kinetic studies for Cd(II) biosorption from treated urban effluents by native grapefruit biomass
(Citrus paradisi L.): The competitive effect of Pb(II), Cu(II) and Ni(II). Chem. Eng. J. 191 (2012), pp.
278–287.
Bhattacharyya, K.G. & Gupta, S.S.: Adsorptive accumulation of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II)
from water on montmorillonite: Influence of acid activation. J. Colloid Interface Sci. 310 (2007),
pp. 411–424.
Bhattacharyya, K.G. & Gupta, S.S.: Adsorption of a few heavy metals on natural and modified kaolinite and
montmorillonite: A review. Adv. Colloid Interface Sci. 140 (2008), pp. 114–131.
Ciesla, P., Kocot, P., Mytych, P. & Stasicka, Z.: Homogeneous photocatalysis by transition metal complexes
in the environment. J. Molec. Catal. A: Chem. 224 (2004), pp. 17–33.
Dhaouadi, A. & Adhoum, N.: Heterogeneous catalytic wet peroxide oxidation of paraquat in the presence of
modified activated carbon. Appl. Catal. B: Environ. 97 (2010), pp. 227–235.
De Leon, M.A., Castiglioni, J., Bussi, J. & Sergio, M.: Catalytic activity of an iron-pillared montmorillonitic
clay mineral in heterogeneous photo-Fenton process. Catal. Today 133 (2008), pp. 600–608.
Ghiselli, G.F., Jardim, W., Litter, M.I. & Mansilla, H.D.: Destruction of EDTA using Fenton and photo-
Fenton-like reactions under UV-A irradiation. J. Photochem. Photobiol. A: Chem. 167 (2004), pp. 59–67.
Gu, X., Evans, L.J. & Barabash, S.J.: Modeling the adsorption of Cd (II), Cu(II), Ni (II), Pb (II) and Zn (II)
onto montmorillonite. Geochim. Cosmochim. Acta 74 ( 2010), pp. 5718–5728.
Hazrat A.: Biodegradation of synthetic dyes—A review. Water Air Soil Pollut. 213 (2010), pp. 251–273.
Herney-Ramirez, J., Vicente, M.A. & Madeira, L.M.: Heterogeneous photo-Fenton oxidation with pillared
clay-based catalysts for wastewater treatment: A review. Appl. Catal. B: Environ. 98 (2010), pp. 10–26.
Hosseini Koupaie, E., Alavi Moghaddam, M.R. & Hashemi, S.H.: Evaluation of integrated anaerobic/aerobic
fixed-bed sequencing batch biofilm reactor for decolorization and biodegradation of azo dye Acid Red
18: Comparison of using two types of packing media. Biores. Technol. 127 (2013), pp. 415–421.
Huanosta-Gutiérrez, T., Dantas, R. F., Ramírez-Zamora, R.M. & Esplugas, S.: Evaluation of copper slag
to catalyze advanced oxidation processes for the removal of phenol in water. J. Hazard. Mater. 213–214
(2012), pp. 325– 330.
Modified montmorillonite in photo-Fenton and adsorption processes 233

Idel-aouad, R., Valiente, M.; Yaacoubi, A., Tanouti, B. & López-Mesas, M.: Rapid decolourization and
mineralization of the azo dye C.I. Acid Red 14 by heterogeneous Fenton reaction. J. Hazard. Mater. 186
(2011), pp. 745–750.
Ijagbemi, C.O., Baek, M.H. & Kim, D.S.: Montmorillonite surface properties and sorption characteristics
for heavy metal removal from aqueous solutions. J. Hazard. Mater. 166 (2009), pp. 538–546.
Khan, R., Bhawana P. & Fulekar, M.H.: Microbial decolorization and degradation of synthetic dyes: A review.
Rev. Environ. Sci. Biotechnol. 12:1 (2013), pp. 75–97.
Klamerth, N., Gernjak, W., Malato, S., Aguera, A. & Lendl, B.: Photo-Fenton decomposition of
chlorfenvinphos: Determination of reaction pathway. Water Res. 43 (2009), pp. 441–449.
Lee, Y. & Lee, W.: Degradation of trichloroethylene by Fe(II) chelated with cross-linked chitosan in a
modified Fenton reaction. J. Hazard. Mater. 178 (2010), pp. 187–193.
Lofrano, G., Meriç, S.V., Belgiorno, N.A. & Napoli, R.M.A.: Fenton and photo-Fenton treatment of
a synthetic tannin used in leather tannery: A multi approach study. Water Sci. Technol. 55 (2007),
pp. 53–61.
Magnoli, A.P., Tallone, L., Rosa, C.A.R., Dalcero, A.M., Chiacchiera, S.M. & Torres Sanchez, R.M.:
Commercial bentonites as detoxifier of broiler feed contaminated with aflatoxin. Appl. Clay Sci. 40
(2008), pp. 63–71.
Masarwa, M., Cohen, H., Meyerstein, D., Hickman, D.L., Bakac, A. & Espenson, J.H.: Reactions of low-
valent transition-metal complexes with hydrogen peroxide. Are they "Fenton-like" or not? The case of
Cu+ and Cr2+ aq. J. Am. Chem. Soc. 110 (1988), pp. 4293–4297.
Meriç, S., Lofrano, G. & Belgiorno, V.: Treatment of reactive dyes and textile finishing wastewater using
Fenton’s oxidation for reuse. Int. J. Environ. Pollut. 23 (2005), pp. 248–258.
Montes, S., Montes-Atenas, G., Salomo, F., Valero, E. & Diaz, O.: On the adsorption mechanisms of copper
ions over modified biomass. Bull. Environ. Contam. Toxicol. 76 (2006), pp. 171–178.
Najjar, W., Azabou, S., Sayadi, S. & Ghorbel, A.: Catalytic wet peroxide photo-oxidation of phenolic olive
oil mill wastewater contaminants: Part I. Reactivity of tyrosol over (Al–Fe)PILC. Appl. Catal. B: Environ.
74 (2007), pp. 11–18.
Neamtu, M., Yediler, A., Siminiceanu, I. & Kettrup, A.: Oxidation of commercial reactive azo dye aqueous
solutions by the photo-Fenton and Fenton-like processes. J. Photochem. Photobiol. 161 (2003), pp. 87–93.
Ohnuki, T., Yoshida, T., Ozaki, T., Samadfam, M., Kozai, N., Yubuta, K., Mitsugashira, T., Kasama, T. &
Francis, A.J.: Interactions of uranium with bacteria and kaolinite clay. Chem. Geol. 220 (2005),
pp. 237–243.
Olivelli, M., Curutchet G. & Torres Sánchez R.: Uranium uptake by montmorillonite-biomass complexes.
Ind. Eng. Chem. Res. 52:6 (2013), pp. 2273–2279.
O’Neill, C., Hawkes, F.R., Hawkes, D.L., Lourenco, N.D., Pinheiro, H.M. & Delee, W.: Color in textile efflu-
ents sources, measurement, discharge consents and simulation: A review. J. Chem. Technol. Biotechnol.
74 (1999), pp. 1009–1018.
Persson, I.: Hydrated metal ions in aqueous solution: How regular are their structures? Pure Appl. Chem. 82
(2010), pp. 1901–1917.
Pignatello, J., Oliveros, E. & MacKay, A.: Advanced oxidation processes for organic contaminant destruction
based on the Fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Tech. 36 (2006), pp. 1–84.
Primo, O., Rivero, M.J. & Ortiz, I.: Photo-Fenton process as an efficient alternative to the treatment of
landfill leachates. J. Hazard. Mater. 153 (2008), pp. 834–842.
Pupo Nogueira, R., Oliveira, M. & Paterlini, W.: Simple and fast spectrophotometric determination of H2 O2
in photo-Fenton reactions using metavanadate. Talanta 66 (2005), pp. 86–91.
Puppo Nogueira, R., Trovó, A., da Silva, M., Villa, R. & Oliveira, M.: Fundamentos e aplicacoes ambientais
dos procesos Fenton e photo-Fenton. Quim. Nova 30 (2007), pp. 400–408.
Reddy, D.H.K., Seshaiah, K., Reddy, A.V.R. & Lee, S.M.: Optimization of Cd(II), Cu(II) and Ni(II)
biosorption by chemically modified Moringa oleifera leaves powder. Carbohyd. Polym. 88 (2012),
pp. 1077–1086.
Schlegel, M.L. & Descostes, M.: Uranium uptake by hectorite and montmorillonite: A solution chemistry
and polarized EXAFS study. Environ. Sci. Technol. 42 (2009), pp. 8593–8598.
Shah, V., Verma, P., Stopka, P., Gabriel, J., Baldrian, P. & Nerud, F.: Decolorization of dyes with
copper(II)/organic acid/hydrogen peroxide systems. Appl. Catal. B: Environ. 46 (2003), pp. 287–292.
Singh, K. & Arora, S.: Removal of synthetic textile dyes from wastewaters: A critical review on present
treatment technologies. Crit. Rev. Environ. Sci. Technol. 41 (2011), pp. 807–878.
Shinde, N.R., Bankar, A.V., Kumar, A.R. & Zinjarde, S.S.: Removal of Ni (II) ions from aqueous solutions
by biosorption onto two strains of Yarrowia lipolytica. J. Environ. Management 102 (2012), pp. 115–124.
234 L.M. Guz et al.

Soon, A.N. & Hameed, B.H.: Heterogeneous catalytic treatment of synthetic dyes in aqueous media using
Fenton and photo-assisted Fenton process. Desalination 269 (2011), pp. 1–16.
Stylidi, M., Kondarides, D.I. & Verykios, X.E.: Pathways of solar light-induced photocatalytic degradation
of azo dyes inaqueous TiO2 suspensions. Appl. Catal. B: Environ. 40:4 (2003), pp. 271–286.
Sykora, J.: Photochemistry of copper complexes and their environmental aspects. Coord. Chem. Rev. 159
(1997), pp. 95–108.
Volesky, B.: Biosorption and me. Water Res. 41 (2007), pp. 4017–4029.
Wu, J., Li, B., Liao, J., Feng, Y., Zhang, D., Zhao, J., Wen, W., Yang, Y. & Liu, N.: Behaviour and analysis of
Cesium adsorption on montmorillonite mineral. J. Environ. Radioactiv. 100 (2009), pp. 914–920.
Yip, A.C.-K., Lam, F.L.-Y. & Hu, X.: Novel bimetallic catalyst for the photo-assisted degradation of Acid
Black 1 over a broad range of pH. Chem. Eng. Sci. 62 (2007), pp. 5150–5153.
Yipmantin, A., Maldonado, H.J., Ly, M., Taulemesse, J.M. & Guibal, E.: Pb(II) and Cd(II) biosorption on
Chondracanthus chamissoi (a red alga). J. Hazard. Mater. 185 (2011), pp. 922–929.
Zhang,S., Liang, S., Wang, X., Long, J., Li, Z. & Wu, L.: Trinuclear iron cluster intercalated montmorillonite
catalyst: Microstructure and photo-Fenton performance. Catal. Today 175 (2011), pp. 362–369.
Zhu, N., Gu, L., Yuan, H., Lou, Z., Wang, L. & Zhang, X.: Degradation pathway of the naphthalene
azo dye intermediate 1-diazo-2- naphthol-4-sulfonic acid using Fenton’s reagent. Water Res. 46 (2012),
pp. 3859–3867.
CHAPTER 14

Photocatalytic degradation of dichlorvos solution using


TiO2 -supported ZSM-11 zeolite

Silvina Gomez, Candelaria Leal Marchena, Luis Pizzio & Liliana Pierella

14.1 INTRODUCTION

Heterogeneous catalysis involving semiconductors has been developed in the last 10–15 years
due to its potential applications to environmental problems (Oancea and Oncescu, 2008). In
recent years, advanced oxidation processes (AOPs) have been proposed as attractive alternatives
for the treatment of contaminated ground-, surface-, and wastewater containing pesticides or
non-biodegradable organic pollutants (Phanikrishna Sharma et al., 2008a). Among the most used
processes are heterogeneous photocatalysis.
The semiconductor TiO2 has been widely utilized as a photocatalyst for inducing a series of
reductive and oxidative reactions on its surface (Chong et al., 2010). TiO2 is one of the most
appropriate material due to its high activity in the photodegradation of organic compounds, low
cost, low toxicity, and chemical stability (Fuchs et al., 2009; Yamaguchi et al., 2009). The anatase
form of TiO2 is more active generally for the degradation of various organic pollutants than
rutile (Mahalakshmi et al., 2009). The efficiency of TiO2 is influenced by many factors, such
as crystalline structure, particle size, and preparation methods (Hsien et al., 2001). However,
poor adsorption and low surface area properties lead to great limitations in exploiting anatase to
the best of its photoefficiency. In addition, technical limitations for adopting this technology are
separation of the catalyst, its reuse and its low quantum efficiency (Phanikrishna Sharma et al.,
2008b). On the other hand, supporting TiO2 is commonly reported to be less photoactive due to
the interaction of TiO2 with the support during the thermal treatments and limitations to mass
transfer (Mohamed et al., 2005).
In the past few years, some efforts have been put in increasing the TiO2 surface area by
dispersing TiO2 nanoparticles on high surface area materials. The supports used included silica
gels, active carbon, zeolites and clays (Anderson and Bard, 1995; 1997; Hermann et al., 1999;
Minero et al., 1992; Torimoto et al., 1997; Xu and Langford, 1997; 1995; Xu et al., 1999;
Yoneyama and Torimoto, 2000). Some of these studies have also included an effort to increase the
adsorption of organic substrates on the catalyst surface for improving the efficiency of catalytic
activity (Hsien et al., 2001).
TiO2 supported on adsorbents provides higher specific surface area and facilitates more effec-
tive adsorption sites than bulk TiO2 (Anderson and Bard, 1995; Takeda et al., 1997; Torimoto
et al., 1997; Xu and Langford, 1995; 1997). Recent research proposed the synthesis of a Fe3+ -
TiO2 photocatalyst supported on zeolite as an alternative for degradation of dyes (Wang et al.,
2011).
Zeolites have been investigated as potential supports for photocatalytic systems. They offer
several distinct advantages over other supports, such as:
• Zeolites have cages and channels in the order of 4–14 Å, which can confine substrate molecules
to enhance the photocatalytic reactivity. TheTiO2 molecule can be thus supported on this zeolite
matrix.
• Zeolites behave as electron donors and as acceptors of moderate strength to the guest species
depending on the adsorption site (Chatti et al., 2007).
235
236 S. Gomez et al.

Organophosphorous pesticides are within the 10 most widely used pesticides all over the world.
They have been used as an alternative to organochlorine compounds for pest control. However, they
are considered extremely toxic compounds acting on acetylcholinesterase activity (Evgenidou
et al., 2006). Their presence as contaminants in aquatic environments may cause serious problems
to human beings and to other organisms. Dichlorvos (DDVP) is an organophosphate compound,
which is used as an agricultural insecticide on crops, stored products and animals. It is also used
as an insecticide for slow release on pest-strips for pest control in homes (Atiqur Rahman and
Muneer, 2005). Thus, the elimination of this pesticide from water is an important research issue.
In the present work, we report the use of zeolite NH4 -ZSM-11 and H-ZSM-11 as the support
for the in situ generation of titania from an alcoholic solution of titanium isopropoxide, followed
by calcination. The resulting catalyst systems were characterized by a series of complementary
techniques: X-ray diffraction spectrometry (XRD), transmittance Fourier transform infrared spec-
troscopy (FTIR) and BET surface area. The supported catalysts were evaluated for their activity
in the photodecomposition of the insecticide dichlorvos (DDVP).

14.2 EXPERIMENTAL

14.2.1 Preparation of zeolite supported TiO2 catalyst


For zeolite synthesis, the following reactants were used: sodium aluminate (NaAlO2 , Jonhson
Matthey Electronics), tetrabutylammonium hydroxide (TBAOH, Fluka), silicic anhydride (Fluka)
and distilled water. The parent Na-ZSM-11 material (Si/Al = 17) was obtained by the hydrother-
mal crystallization method (Anunziata and Pierella, 1993). Aqueous solution of sodium aluminate
was introduced into a silica solution, previously prepared by partial dissolution of silicic anhy-
dride on tetrabutylammonium hydroxide (TBAOH). The obtained gel reached pH > 9 and was
maintained at 120–160◦ C for 12–16 days under self-generated pressure on an autoclave. Then,
reaction products were extracted, washed and dried at 110◦ C for 12 h. The structure directing
agent (TBAOH) was desorbed under N2 atmosphere (20 mL/min) at programmed temperature
(10◦ C/min) from 110 to 520◦ C, and then calcined in air at 520◦ C for 12 h to obtain Na-
ZSM-11. The ammonium form of the zeolite (NH4 -ZSM-11) was prepared by ion exchange
of the as-prepared Na-zeolite form with 1 M ammonium chloride solution at 80◦ C for 40 h.
Finally, NH4 -zeolite was dried at 110◦ C, treated in a nitrogen flow at 500◦ C during 8 h and
then calcined in air at the same temperature for 10 h obtaining the HZSM-11 (Anunziata and
Pierella, 1993).
The supported catalysts were prepared by taking the appropriate amount of titanium (IV)
isopropoxide (as source of titanium) and zeolite (HZSM-11 or NH4 -ZSM-11) in ethanol. This
mixture was mechanically stirred for 4 h at ambient temperature. Then, the solvent was removed
by rotary evaporation a 50◦ C. The amount of titanium isopropoxide was varied with the purpose
of generating in situ TiO2 concentrations of 10, 20, 30 and 50% by weight in the final solid. The
mixture was then dried at 110◦ C and calcined in air at 450◦ C.
The TiO2 /HZSM-11(30%) sample was also calcined at 600 and 800◦ C to study the effect of
the calcination temperature.
A physical mixture of P25/zeolite (30/70 weight %) was prepared for comparison. This mixture
was then calcined at 450◦ C for 8 h.
The commercially available TiO2 (Degussa P-25 with 70wt% anatase and 30wt% rutile, surface
area 50 m2 /g and particle size 25 nm) obtained from Degussa Chemical, Germany (now Evonik),
was used as received.

14.2.2 Characterization of the photocatalysts


The powder XRD diffraction patterns of the materials were collected on a PANalytical X’pert
PRO diffractometer equipped with Cu Kα (1.54 Å) in the range of 2θ from 5–60◦ , in steps of
Photocatalytic degradation of dichlorvos solution 237

0.05◦ , with a count time of 2 s at each point. The average crystallite sizes of TiO2 crystals over
the zeolite matrix were estimated using the XRD data and the Scherrer equation:
d = (k × λ) ÷ (b ×cos θ ) (14.1)

where d is the crystallite diameter in nm, k is the shape constant (0.9), λ is the wavelength in nm,
θ is the Bragg angle in degrees and b is the observed peak width at half-maximum peak height in
rad (Luísa et al., 2011).
BET surface area determinations were carried out with a Micromeritics ASAP 2000 equipment.
Infrared (IR) studies of TiO2 /zeolites were performed on a JASCO 5300 FTIR spectrometer.
The spectra in the lattice vibration region were performed using KBr 0.05% wafer technique and
they were carried out from 1800 to 400 cm−1 in 16 consecutive registers of 4 cm−1 resolution
each one.

14.2.3 Photocatalytic experiments and analyses


Photocatalytic degradation of DDVP (Pestanal, Fluka) was performed in aqueous medium in
a batch reactor. A cylindrical Pyrex glass photochemical reactor of 18 cm × 8 cm (height ×
diameter), provided with a water circulation arrangement to maintain the temperature in the range
25–30◦ C, was used in all the experiments. The irradiation was carried out using a 125 W high-
pressure mercury lamp placed inside a Pyrex glass jacket, thermostatted by water circulation,
and immersed in the pesticide solution contained in the reactor. The catalyst was maintained
in suspension by stirring, and air was continuously bubbled. Previously, the DDVP solution
(400 mL, 1 × 10−4 M), containing 100 mg of catalyst was magnetically stirred in the absence of
light for 30 min. Aliquots were withdrawn at specific time intervals and analyzed after filtration to
remove the catalyst. The variation of the pesticide concentration as a function of the reaction time
was determined by a UV-visible Shimadzu spectrophotometer of double beam, measuring the
absorbance at 210 nm (Atiqur Rahman and Muneer, 2005). In some cases, a HPLC Perkin Elmer
equipment, with a 250 mm × 4.6 mm C18 reverse phase column, and a mixture of water (50%)
with acetonitrile (50%) as solvent, 1 mL/min flow rate, was employed. An ion chromatograph
(Metrohm – 761 compact IC) equipped with an electroconductivity detector was used to measure
sulfate ions. A chloride electrode (Phoenix Clo 1508-003B) was used to follow chloride ions.

14.3 RESULTS AND DISCUSSION

14.3.1 Characterization of TiO2 /zeolite catalysts


14.3.1.1 XRD analysis
In order to confirm the crystalline structure of the TiO2 /zeolite catalysts, an XRD study was
carried out. Figure 14.1 shows the XRD patterns of H-ZSM-11 and of different titania-supported
catalysts. For comparative purposes, the diffractogram of commercial P-25 was also included.
Figure 14.1 shows the characteristic signals of the parent HZSM-11 catalyst at 2θ angles of
23–24◦ and 7–9◦ (Renzini et al., 2009).
TiO2 /HZSM-11 prepared in this study showed diffraction peaks at 25.43, 48.16, 54.9 and 55.4◦ ,
which were assigned to the characteristic reflections from the (101), (200) and (211) and (106)
planes of anatase, respectively (Porkodi and Arokiamary, 2007). In the diffraction pattern of
TiO2 /HZSM-11(10%), the peak at (101) is barely seen, but the other three peaks were not clearly
observed due to the low amount of TiO2 .
The peak intensity at 2θ = 25.43◦ increased with the amount of TiO2 loaded on HZSM-11.
A similar XRD pattern was obtained for TiO2 /zeolite samples reported previously (Chen et al.,
1999; Mahalakshmi et al., 2009; Petkowicz et al., 2009; Yamaguchi et al., 2009). In hand with the
increase in the TiO2 load, a small decrease in the intensity of the diffraction peak of HZSM-11 is
observed. However, this decrease can be attributed to a dilution effect of the zeolite in the catalyst.
238 S. Gomez et al.

Figure 14.1. XRD pattern of (a) HZSM-11, (b) TiO2 /HZSM-11 (10%), (c) TiO2 /HZSM-11 (20%), (d)
TiO2 /HZSM-11(30%), (e) TiO2 /HZSM-11(50%) and (f) P-25.

Figure 14.2. XRD pattern of the (a) zeolite NH4 -ZSM-11, (b) TiO2 /NH4 ZSM-11 (10%), (c) TiO2 /
NH4 ZSM-11(20%), (d) TiO2 / NH4 ZSM-11(30%), (e) TiO2 /NH4 ZSM-11(50%) and (f) P-25
Degussa.

On the other hand, no signals corresponding to the rutile phase (2θ = 27.4◦ ) could be detected
in the XRD patterns of all TiO2 /HZSM-11 catalysts used in the present study.
The XRD patterns of NH4 -ZSM-11, TiO2 Degussa and NH4 -ZSM-11 supported TiO2 catalysts
are shown in Figure 14.2. As can be seen, the results obtained were very similar to the HZSM-11
zeolite (Fig. 14.1).
Photocatalytic degradation of dichlorvos solution 239

Figure 14.3. XRD pattern of the TiO2 /HZSM-11 materials at TiO2 30 wt% obtained for calcination
temperatures of (a) 450◦ C, (b) 600◦ C and (c) 800◦ C.

Table 14.1. Crystallite size of the two series of


zeolite samples.

Sample d (nm)

HZSM-11 series
TiO2 /HZSM11(10%) 12.04
TiO2 /HZSM11(20%) 13.13
TiO2 /HZSM11(30%) 18.86
TiO2 /HZSM11(50%) 22.98
NH4 ZSM-11 series
TiO2 /NH4 ZSM11(10%) 13.12
TiO2 /NH4 ZSM11(20%) 14.13
TiO2 /NH4 ZSM11(30%) 15.98
TiO2 /NH4 ZSM11(50%) 22.98

The calcination temperature effect over TiO2 /HZSM-11(30%) photocatalysts is shown in Figure
14.3. The diffraction patterns of the materials did not show the presence of the rutile phase.
Therefore, it can be suggested that the anatase particles are favorably stabilized on the surface of
the zeolite matrix.
The calcination temperature profoundly influenced the crystallite size of the TiO2 of the sup-
ported catalyst. As result of the increment of the calcination temperature from 450 to 800◦ C, an
increase in the size of the anatase crystals was observed. In Figure 14.3, it can be seen that the
XRD peaks become gradually sharper with the increasing temperature, indicating that the crystals
grew larger in size (Porkodi and Arokiamary, 2007).
The TiO2 crystallite size values estimated during this work are presented in Table 14.1. All the
TiO2 crystals supported in the zeolite matrix have dimensions on the nanometer range, varying
from 12 to 23 nm.
The results as a whole indicate that the TiO2 particle size increased with the increasing TiO2
loading, possibly due to aggregation of the TiO2 particles on the surface of the zeolite.
240 S. Gomez et al.

Figure 14.4. FTIR spectra of (a) HZSM-11, (b) TiO2 /HZSM-11(10%), (c) TiO2 /HZSM-11(20%), (d)
TiO2 /HZSM-11(30%), (e) TiO2 /HZSM-11(50%) and (f) P-25.

14.3.1.2 FTIR spectra


Figure 14.4 shows the FTIR spectra of pure TiO2 , pure zeolite and TiO2 /HZSM-11 (wt%) in the
1800–400 cm−1 range. All the zeolitic materials present vibrations assigned to the internal bonds
of a TO4 tetrahedral structure (T = Si or Al) that are insensitive to changes in the zeolite structure:
bands at 1250–950, 850–700 cm−1 and from 500 to 420 cm−1 are attributed to asymmetrical
stretching, to the symmetrical stretching and to (O–T–O) deformation, respectively.
Furthermore, vibrations assigned to the external bonds of the tetrahedral TO4 , which are sensi-
tive to changes in the structure, can also be observed. These vibrations, in the region between 650
and 500 cm−1 , are attributed to those of double rings consisting of five atoms. A shoulder between
1050 and 950 cm−1 is assigned to the asymmetrical stretching of T–O–T bonds (Anunziata et al.,
2004).
No band was detected in the region near 960 cm−1 , which could be assigned to the antisymmetric
stretching vibration of the Ti–O–Si bonds (Kim and Yoon, 2001; Petkowicz et al., 2010; Wang
et al., 2008). Therefore, the replacement of the tetrahedral Si sites with Ti during the preparation
did not take place. This suggests that TiO2 was deposited on the surface of the zeolites.
Figure 14.5 shows the FTIR spectra of the zeolite matrix NH4 -ZSM-11, the supported catalyst
TiO2 /NH4 ZSM-11(wt%) and P25. The spectra were very similar to that obtained for the HZSM-11
samples. The conclusions regarding the Figure are the same.

14.3.1.3 BET surface area


The surface areas (SBET ) of the synthesized materials were determined from the adsorption-
desorption of nitrogen isotherms using the method of Brunauer-Emmett-Teller. The
adsorption-desorption isotherms of TiO2 /HZSM-11(wt%) and TiO2 /NH4 ZSM-11(wt%) exhibit
characteristics similar to HZSM-11 and NH4 -ZSM-11, respectively. According to IUPAC
classification, they are Type I isotherms, characteristic of microporous solids having relatively
small external surfaces.
The surface area of the zeolite matrices and supported catalysts decrease when increasing the
TiO2 loading in the matrix (Table 14.2). This may be due to aggregation of TiO2 particles on the
surface of the materials and blocking of the pores of the zeolite.
Photocatalytic degradation of dichlorvos solution 241

Figure 14.5. FTIR of the (a) zeolite NH4 ZSM-11, (b) TiO2 /NH4 ZSM-11(3%), (c) TiO2 /NH4 ZSM-
11(10%), (d) TiO2 /NH4 ZSM-11(20%), (e) TiO2 /NH4 ZSM-11(30%), (f) TiO2 /NH4 ZSM-
11(50%) and (g) P-25 Degussa.

Table 14.2. BET surface area of materials.

Sample SBET Sample SBET


(m2 g−1 ) (m2 g−1 )

HZSM-11 series NH4 ZSM-11 series


HZSM-11 392 NH4 -ZSM-11 382
TiO2 /HZSM11(10%) 365 TiO2 /NH4 ZSM11(10%) 319
TiO2 /HZSM11(20%) 328 TiO2 /NH4 ZSM11(20%) 305
TiO2 /HZSM11(30%) 309 TiO2 /NH4 ZSM11(30%) 302
TiO2 /HZSM11(50%) 255 TiO2 /NH4 ZSM11(50%) 268

On the other hand, the surface area of the TiO2 /HZSM-11(30%) calcined at different tem-
perature decreases with the increase of the calcination temperature, as shown in Figure 14.6.
The discrepancy on the surface area values can be explained by the TiO2 agglomeration/growth
phenomena during the calcination treatments.

14.3.2 Photocatalytic evaluation


14.3.2.1 Preliminary studies
Previously to photocatalytic experiments, the DDVP solution (400 mL, 1 × 10−4 M) containing
100 mg of catalyst was magnetically stirred in the absence of light to study the adsorption on
the surface of the materials. The maximum adsorption was reached within 30 min. These results
show that it is necessary at least 30 min to ensure that the adsorption-desorption equilibrium of
DDVP on the surface of the materials was attained.
The photolysis results (Fig. 14.7) showed only 2–5% degradation of the DDVP. In the
presence of HZSM-11 or NH4-ZSM-11 as photocatalyst, no significant degradation was
observed, confirming that the zeolite matrix is not active by itself for the degradation of DDVP
(Fig. 14.7).
242 S. Gomez et al.

Figure 14.6. BET surface area vs. calcination temperature in the TiO2 /HZSM-11(30%) catalyst.

Figure 14.7. Preliminary studies: Photolysis (), Photocatalytic degradation with HZSM-11 (♦) and NH4 -
ZSM-11 ().

14.3.2.2 Effect of TiO2 content on TiO2 /HZSM-11 and TiO2 /NH4 -ZSM-11 samples
The effect of the TiO2 content (10, 20, 30 and 50 wt%) over the zeolite matrices in the photocat-
alytic activity was investigated and the results are provided in Figure 14.8. The studies were carry
out using 100 mg of catalyst in 400 mL of 1 × 10−4 M DDVP aqueous solution, and the degra-
dation percentage was calculated as X = (C0 − C) × 100/C0 , where C0 is the original DDVP
concentration, and C is the remaining DDVP concentration in solution at specific time intervals.
Photocatalytic degradation of dichlorvos solution 243

Figure 14.8. Degradation of DDVP at 240 min of reaction as a function of the TiO2 loading in HZSM-11
() and NH4-ZSM-11 (♦) samples.

Figure 14.9. Concentration of [Cl− ] in solution in function of the TiO2 content on HZSM-11 () and
NH4 -ZSM-11 (♦) at 240 minutes of reaction.

The increment of the TiO2 content in the zeolite matrices from 10 to 30 wt% produces a continuous
increment of the photocatalytic activity. The decline in the degradation for samples with loadings
higher than 30% TiO2 /zeolite could be due to the presence of aggregates of TiO2 particles of
large size over the zeolite matrix surface.
Figure 14.9 shows that an increase in the DDVP degradation generates an increase in the
concentration of chloride ions in solution, indicating the dechlorination of DDVP.
The catalyst with the best performance for the DDVP degradation (TiO2 /HZSM-11(30%)) was
chosen for further studies.
244 S. Gomez et al.

Figure 14.10. DDVP degradation and Cl− , PO3−


4 release employing TiO2 /HZSM-11(30%) as
photocatalyst.

The DDVP degradation was monitoring following, Cl− and PO3− 4 as mineralization products
as a function of the irradiation time (Fig. 14.10).
After 240 min of irradiation, the amount of DDVP degraded was close to 90%. It can be
observed that at the same irradiation time, the molar ratio between the amount of Cl− released
and the amount of DDVP degraded is close to two. These results are in agreement with previous
reports that indicate that the cleavage of a Cl-aryl bond, releasing inorganic chloride to the reacting
medium, is the first step during the photocatalytic degradation of DDVP (Evgenidou et al., 2006).
On the other hand, the amount of PO4 3-anions released was only 15% of the total amount
of phosphate present in the original DDVP solution, which indicates the formation of phosphate
containing intermediates (Evgenidou et al., 2005). Thus, it is not possible to reach the total
mineralization of the pollutant under these conditions.

14.3.2.3 Effect of the preparation of the catalyst and role of the support
According to Chatti et al. (2007), the TiO2 incorporation mode played an important role in the
photocatalytic activity of the synthesized photocatalysts. Figure 14.11 shows the photocatalytic
activity of TiO2 /HZSM-11(30%) and a physical mixture of TiO2 and HZSM-11 containing the
same amount of titania that the sample prepared by the impregnation method. We can see that
catalyst synthesized using an alcoholic solution of titanium isopropoxide shows the best photo-
catalytic activity. This may be attributed to a better dispersion and interaction of TiO2 particles
over the zeolite matrix.
On the other hand, to understand the role of the support during the photocatalytic degradation
(i.e., beneficial or detrimental effect), comparative studies were carried out between TiO2 /HZSM-
11(30%) (obtained by impregnation), P25 and TiO2 prepared in a similar way as that those
supported on ZSM zeolite. The reactions with P25 and TiO2 were carried out with 30 mg of the
catalyst, equivalent weight at 30% in the zeolite matrix. In Figure 14.11, it is observed that the
TiO2 supported system shows similar degradation compared to P25 at the end of the reaction (240
minutes). This may be caused by the synergistic effect resulting from the adsorption of DDVP over
the material, which can facilitate the degradation without affecting the photocatalytic properties
Photocatalytic degradation of dichlorvos solution 245

Figure 14.11. Comparative performance of the synthesized photocatalysts: physical mixture (P25/zeolite
HZSM-11), TiO2 /HZSM-11 (30%) (obtained by impregnation), P25 and TiO2 sol-gel (pre-
pared in a similar way as that those supported on ZSM zeolite) at 240 minutes of the
reaction.

of TiO2 . The delocalization of electrons at HZSM-11 framework and the adsorption properties of
the support enhance the degradation. In addition, the dispersion of TiO2 over the zeolite material
avoids particle–particle aggregation and light scattering by TiO2 . The good dispersion leads to
facilitate the presence of holes near the adsorbed DDVP molecules, resulting in faster degradation
rates.

14.3.2.4 Effect of catalyst amount


The effect of the catalyst concentration on the degradation of DDVP was investigated employing
different concentrations of TiO2 /HZSM-11(30%), in the range 50–1000 mg L−1 , keeping all
other experimental parameters constant (Fig. 14.12). The DDVP degradation increases with the
increment of the catalyst concentration from 50 to 250 mg L−1 . However, further increments in
catalyst concentration lead to a continuous decrease on the amount of DDVP degraded. When the
TiO2 /HZSM-11(30%) concentration is high (above 250 mg L−1 ), a screening effect is produced.
The degradation diminishes due the excessive opacity of the suspension, which prevents a good
illumination of the catalyst (Chen et al., 1999). Further, at higher catalyst loading it is difficult
to maintain homogeneous the suspension due to the agglomeration of particles, which decreases
the number of active sites (Mahalakshmi et al., 2009).

14.3.2.5 Effect of the calcination temperature


As mentioned in Section 14.3.1.1., higher calcination temperatures might cause large TiO2 crys-
tallites to form on the surface of the zeolite support. Therefore, it is to be expected that catalysts
treated at lower calcination temperature might display better photocatalytic degradation per-
formance. As shown in Figure 14.13, photodegradation efficiency increased with decreasing
calcination temperature, and the optimum temperature was found to be 450◦ C. The sharp decrease
on photocatalytic degradation for the catalyst treated at 600 and 800◦ C was due to an overgrowth
of the TiO2 crystallite size.
246 S. Gomez et al.

Figure 14.12. Effect of TiO2 /HZSM-11(30%) catalyst amount for DDVP degradation activity.

Figure 14.13. Effect of calcination temperature of TiO2 /HZSM-11(30%) for DDVP degradation activity.

14.3.2.6 Effect of initial pH value


According to Mahalakshmi et al. (2009), pH is an important variable in aqueous phase mediated
photocatalytic reactions. The pH of the solution influences adsorption and dissociation of sub-
strate, catalyst surface charge, oxidation potential of the valence band and other physicochemical
properties of the system. The effect of pH on the photocatalytic degradation was studied varying
the initial pH of the DDVP solution and keeping all other experimental conditions constant. The
initial solution pH in the reactions evaluated in previous paragraphs was 5. In order to compare
the standard reaction with the initial pH, two experiments were carried out at pH 2 and at pH 10,
adjusted with HCl and NaOH, respectively. The results at 240 min are depicted in Figure 14.14.
As it can be seen, the DDVP degradation decreases when pH was adjusted to 2, and remains
Photocatalytic degradation of dichlorvos solution 247

Figure 14.14. Effect of pH for DDVP degradation activity with TiO2 /HZSM-11(30%) () and
TiO2 /NH4 ZSM-11(30%) (♦) at 240 minutes of reaction.

practically constant when it was equal to 10. This can be observed for the two series HZSM-11
and NH4 ZSM-11 of supported catalysts.
The effect of pH on the photocatalytic reaction is generally attributed to the surface charge of
TiO2 (point of zero charge, pz, of TiO2 is 5) and its relation with the ionic form of the organic com-
pound (anionic or cationic) (Fernández-Ibáñez et al., 2003). Electrostatic attraction or repulsion
between the surface of the catalyst and the organic molecule is taking place and, consequently,
it enhances or inhibits, respectively, the photodegradation rate. According to Evgenidou et al.
(2005), the reaction at alkaline pH can be ascribed to the high hydroxylation of the surface of the
catalyst due to the presence of a large quantity of OH− ions. Consequently, higher concentrations
of OH− species are formed and the overall rate is enhanced. However, in Figure 14.14, it can be
seen that for both supported catalysts reactions at pH 10 and pH 5 present similar degradation
extent.

14.3.2.7 Effect of adding H2 O2 to the photodegradation of DDVP


It has been pointed out that H2 O2 addition to a heterogeneous photocatalytic degradation process
increases the efficiency of the process (Naman et al., 2002), because H2 O2 is an efficient elec-
tron acceptor that increases the generation of OH radicals when reacting with conduction band
electrons on the surface of the semiconductor:

H2 O2 + e− −
CB → OH + HO

(14.2)

Figures 14.15 and 14.16 show the comparison of DDVP photodegradation in the presence
of TiO2 /HZSM-11(30%) and TiO2 /NH4 ZSM-11(30%), respectively, with and without H2 O2
(10 mM) addition and keeping all other experimental conditions constant. As can be seen, H2 O2
addition improved DDVP photodegradation, and a complete DDVP degradation was reached after
3 h of irradiation with TiO2 /HZSM-11(30%) and a 98% degradation percentage after 240 minutes
of irradiation with TiO2 /NH4 ZSM-11(30%).
248 S. Gomez et al.

Figure 14.15. Effect of addition of H2 O2 for DDVP degradation activity with TiO2 /HZSM-11(30%).

Figure 14.16. Effect of H2 O2 addition in the DDVP degradation activity with TiO2 /NH4 ZSM-11(30%).

14.3.3 Photocatalyst recycling studies


At the end of the first catalytic cycle, the catalyst was filtered, washed, oven dried at 70◦ C and
then reused without calcination treatments. The same procedure was employed after a second
cycle. From the comparison of DDVP degradation degrees reached at the end of the 1st, 2nd
and 3rd cycle, a slight decrease in the degradation rate was noticed (Fig. 14.17). This decrease
was probably due to the accumulation of organic intermediates in the cavities and on the surface
Photocatalytic degradation of dichlorvos solution 249

Figure 14.17. DDVP degradation degree as a function of the number of cycles of reuse for the TiO2 /
HZSM-11(30%) catalyst.

Figure 14.18. XRD of TiO2 / HZSM-11(30%). (a) Fresh, (b) used – without regeneration (after cycle 3rd)
and (c) used – with calcination (after the 4th cycle).

of the photocatalyst, affecting DDVP adsorption, and reducing the activity of the material. The
catalyst recovered after the 3rd cycle was calcined at 450◦ C for 8 h and reused. As it can be seen,
the original activity was restored (Fig. 14.17, 4th cycle).
Hence, calcination of the reused catalyst is necessary in order to regenerate the activity. Fur-
thermore, this is substantiated by comparing the surface characterization studies of fresh and used
catalysts (before and after the 4th cycle) using XRD (Fig. 14.18). The XRD pattern showed that
the sample was intact even after the 4th cycle, comparing the TiO2 characteristic peaks in the
250 S. Gomez et al.

Figure 14.19. Temporal evolution of DDVP after photocatalysis with TiO2 /HZSM-11(30%) (♦) and
leaching test () with the same filtrate catalyst at the 60 minutes.

fresh and used catalyst. Thus, it proves that the catalyst is highly stable and is reusable for several
cycles without losing its original activity.
It is worth to be noted that no leaching of TiO2 from TiO2 /ZSM-11 samples occurred during the
reaction. Additionally, no significant conversion of dichlorvos was observed when the reaction
was carried out by dissolving DDVP in the liquid filtrate in the absence of TiO2 /HZSM-11(30%)
(Fig. 14.19). These observations strongly suggest that the reaction proceeds heterogeneously over
the catalyst.

14.4 CONCLUSIONS

Catalyst based on titania supported on ZSM-11 and NH4 ZSM-11 zeolite were prepared by taking
an amount of titanium (IV) isopropoxide and of zeolite in ethanol. The XRD patterns of the
materials did not show the presence of the rutile phase, whereas the anatase phase was the only
phase observed independently of calcination temperatures and TiO2 content. It can be suggested
that the anatase particles are stabilized in the surface of the zeolite matrix.
The calcination temperature is an important variable of the catalyst preparation and influences
the photocatalytic activity. The higher the calcination temperature, the higher the crystallite size,
with a lower photocatalytic activity.
The TiO2 /HZSM-11 and TiO2 /NH4 ZSM-11 catalysts prepared in this study were suitable for
the degradation of the insecticide DDVP in water, resulting in degradation percentages close
to that of commercial P25. The advantage of these supported catalysts, in comparison with the
commercial, is mainly the easy separation from the aqueous system and the possibility of reuse.
The results of the present investigation conclude that the best catalyst was TiO2 /ZSM-11(30%)
prepared by the impregnation method. Several variables were studied, and it was observed that the
best operating conditions were: pH 5, 450◦ C as the calcination temperature of the photocatalyst
and catalyst/solution ratio equal to 250 mg L−1 . The addition of hydrogen peroxide increased the
photocatalytic degradation of dichlorvos.
After regenerating the materials not a significant change in the structure and activity was
observed, indicating that they can be reused without noticeable activity loss during at least four
cycles.
Photocatalytic degradation of dichlorvos solution 251

ACKNOWLEDGEMENTS

This project was partially supported by UTN-PID 25E129. We thank CONICET: L.B. Pierella
(researcher), L. Pizzio (researcher), S. Gomez (doctoral fellowship) and C. Leal Marchena
(doctoral fellowship).

REFERENCES

Anderson, C. & Bard, A.J.: An improved photocatalyst of TiO2 /SiO2 prepared by a sol-gel synthesis. J. Phys.
Chem. 99 (1995), pp. 9882–9885.
Anderson, C. & Bard, A.J.: Improved photocatalytic activity and characterization of mixed TiO2 /SiO2 and
TiO2 /Al2O3 materials. J. Phys. Chem. B 101 (1997), pp. 2611–2616.
Anunziata, O. & Pierella, L.: Nature of the active sites in H-ZSM-11zeolite modified with Zn(+2) and Ga(+3 ).
Catal. Lett. 19: 2–3 (1993), pp. 143–151.
Anunziata, O., Beltramone, A., Juric, Z., Pierella, L. & Requejo, F.: Fe-containing ZSM-11 zeolites as
active catalyst for SCR of NOx Part I. Synthesis, characterization by XRD, BET and FTIR and catalytic
properties. Appl. Catal. A: Gen. 264 (2004), pp. 93–101.
Atiqur Rahman, M. & Muneer, M.: Photocatalysed degradation of two selected pesticide derivatives,
dichlorvos and phosphamidon, in aqueous suspensions of titanium dioxide. Desalination 181 (2005),
pp. 161–172.
Chatti, R., Rayalu, S., Dubey, N., Labhsetwar, N. & Devotta, S.: Solar-based photoreduction of methyl
orange using zeolite supported photocatalytic materials. Sol. Ener. Mat. Sol. C 91 (2007), pp. 180–190.
Chen, H., Matsumoto, A., Nishimiya, N. & Tsutsumi, K.: Preparation and characterization of TiO2
incorporated Y-zeolite. Colloid. Surface A 157 (1999), pp. 295–305.
Chong, M.N., Jin, B., Chow, C. & Saint, C.: Recent developments in photocatalytic water treatment
technology: A review. Water Res. 44 (2010), pp. 2997–3027.
Evgenidou, E., Fytianos, K. & Poulios, I.: Semiconductor-sensitized photodegradation of dichlorvos in water
using TiO2 and ZnO as catalysts. Appl. Catal. B: Environ. 59 (2005), pp. 81–89.
Evgenidou, E., Konstantinou, I., Fytianos, K. & Albanis, T.: Study of the removal of dichlorvos and
dimethoate in a titanium dioxide mediated photocatalytic process through the examination of intermediates
and the reaction mechanism. J. Hazard. Mater. B137 (2006), pp. 1056–1064.
Fernández-Ibáñez, P., Blanco, J., Malato, S. & De las Nieves, F.P.: Application of the colloidal stability of
TiO2 particles for recovery and reuse in solar photocatalysis. Water Res. 37 (2003), pp. 3180–3188.
Fuchs, V., Méndez, L., Blanco, M. & Pizzio, L.: Mesoporous titania directly modified with tungstophos-
phoric acid: Synthesis, characterization and catalytic evaluation. Appl. Catal. A: Gen. 358 (2009),
pp. 73–78.
Hermann, J., Matos, J., Disdier, J., Guillard, C., Laine, J., Malato, S. & Blanco, J.: Solar photocatalytic
degradation of 4-chlorophenol using the synergistic effect between titania and activated carbon in aqueous
suspension Catal. Today 54 (1999), pp. 255–265.
Hsien, Y-H., Chang, C-F., Chen, Y-H. & Cheng, S.: Photodegradation of aromatic pollutants in water over
TiO2 supported on molecular sieves. Appl. Catal. B: Environ. 31 (2001), pp. 241–249.
Kim, Y. & Yoon, M.: TiO2 /Y-Zeolite encapsulating intramolecular charge transfer molecules: A new photo-
catalyst for photoreduction of methyl orange in aqueous medium. J. Mol. Catal. A: Chem. 168 (2001),
pp. 257–263.
Luísa, A.M., Nevesb, M.C., Mendonc, M.H. & Monteiroc, O.C.: Influence of calcination parameters on the
TiO2 photocatalytic properties. Mater. Chem. Phys. 125 (2011), pp. 20–25.
Mahalakshmi, M., Vishnu Priya, S., Arabindoo, B., Palanichamy, M. & Murugesan, V.: Photocatalytic
degradation of aqueous propoxur solution using TiO2 and Hβ zeolite-supported TiO2 . J. Hazard. Mater.
161 (2009), pp. 336–343.
Minero, C., Catozzo, F. & Pelizzetti, E.: Role of adsorption in photocatalyzed reactions of organic molecules
in aqueous Ti02 suspensions. Langmuir 8 (1992), pp. 481–486.
Mohamed, R.M, Ismail, A.A, Othman, I. & Ibrahim, I.A: Preparation of TiO2 -ZSM-5 zeolite for
photodegradation of EDTA. J. Mol. Catal. A: Chem. 238 (2005), pp. 151–157.
Naman, S.A, Khammas, Z.A-A. & Hussein, F.M.: Photo-oxidative degradation of insecticide dichlorovos
by a combined semiconductors and organic sensitizers in aqueous media. J. Photochem. Photobiol. A:
Chem. 153 (2002), pp. 229–236.
252 S. Gomez et al.

Oancea, P. & Oncescu, T.: The photocatalytic degradation of dichlorvos under solar irradiation. J. Photochem.
Photobiol. A: Chem. 199 (2008), pp. 8–13.
Phanikrishna Sharma, M.V, Lalitha K., Durgakumari, V. & Subrahmanyam, M.: Solar photocatalytic
mineralization of isoproturon over TiO2 /HY composite systems. Sol. Ener. Mat. Sol. C. 92 (2008a),
pp. 332–342.
Phanikrishna Sharma, M.V, Durga Kumari, V. & Subrahmanyam, M.: TiO2 supported over SBA-15:
An efficient photocatalyst for the pesticide degradation using solar light. Chemosphere 73 (2008b),
pp. 1562–1569.
Petkowicz, D.I., Brambilla, R., Radtke, C., Silva da Silva, C.D, N. da Rocha, Z., Pergher, S.B.C. & Dos
Santos, J.H.Z.: Photodegradation of methylene blue by in situ generated titania supported on a NaA
zeolite. Appl. Catal. A: Gen. 357 (2009), pp. 125–134.
Petkowicz, D.I., Pergher, S.B.C, Silva da Silva, C.D., Novais da Rocha, Z. & Dos Santos, J.H.Z.: Catalytic
photodegradation of dyes by in situ zeolite supported titania. Chem. Eng. J. 158 (2010), pp. 505–512.
Porkodi, K. & Arokiamary, S.D.: Synthesis and spectroscopic characterization of nanostructured anatase
titania: A photocatalyst. Mater. Charact. 58 (2007), pp. 495–503.
Renzini, M.S., Sedrán, U. & Pierella, L.B.: H-ZSM-11 and Zn-ZSM-11 zeolites and their applications in the
catalytic transformation of LDPE. J. Anal. Appl. Pyrolysis 86 (2009), pp. 215–220.
Takeda, N., Ohtani, M., Torimoto, T., Kuwabata, S. & Yoneyama, H.: Evaluation of diffusibility of adsorbed
propionaldehyde on titanium dioxide-loaded adsorbent photocatalyst films from its photodecomposition
rate. J. Phys. Chem. 101 (1997), pp. 2644–2649.
Torimoto, T., Okawa, Y., Takeda, N. & Yoneyama, H.: Effect of activated carbon content in TiO2 -1oaded
activated carbon on photodegradetion behaviors of dichloromethane. J. Photochem. Photobiol. A: Chem.
103 (1997), pp. 153–157.
Wang, C.-C., Lee, C.-K., Lyu, M.-D. & Juang, L.-C.: Photocatalytic degradation of C.I. Basic Violet 10
using TiO2 catalysts supported by Y zeolite: an investigation of the effects of operational parameters.
Dyes Pigments 76 (2008), pp. 817–824.
Wang, C.-C, Shi, H. & Li, Y.: Synthesis and characteristics of natural zeolite supported Fe3+ -TiO2
Photocatalysts. Appl. Surf. Sci. 257 (2011), pp. 6873–6877.
Xu, Y. & Langford, C.H.: Enhanced photoactivity of a titanium(1V) oxide supported on ZSMS and zeolite
A at low coverage. J. Phys. Chem. 99 (1995), pp. 11,501–11,507.
Xu, Y. & Langford, C.H.: Photoactivity of titanium dioxide supported on MCM41, Zeolite X, and zeolite Y.
J. Phys. Chem. B 101 (1997), pp. 3115–3121.
Xu, Y., Zheng, W. & Liu, W.: Enhanced photocatalytic activity of supported TiO2 : Dispersing effect of SiO2 .
J. Photochem. Photobiol. A: Chem. 122 (1999), pp. 57–60.
Yamaguchi, S., Fukura, T., Imai, Y., Yamaura, H. & Yahiro, H.: Photocatalytic activities for partial oxidation
of -methylstyrene over zeolite-supported titanium dioxide and the influence of water addition to reaction
solvent. Electrochim. Acta 55 (2009), pp. 7745–7750.
Yoneyama, H. & Torimoto, T.: Titanium dioxide/adsorbent hybrid photocatalysts for photodestruction of
organic substances of dilute concentrations. Catal. Today 58 (2000), pp 133–140.
CHAPTER 15

Water disinfection with UVC and/or chemical inactivation.


Mechanistic differences, implications and consequences

Marina Flores, Rodolfo Brandi, Alberto Cassano & Marisol Labas

15.1 INTRODUCTION

Human society requires water for drinking, sanitation, cleaning, production of food and energy,
and support of commercial and industrial activities.
Water in nature can contain a variety of contaminants such as minerals, salts, heavy metals,
organic compounds, radioactive residues and living materials, for example parasites, fungi, and
bacteria (US EPA, 2003). In rural and urban areas of low-income countries, millions of the most
vulnerable people lack access to improved water, sanitation and hygiene (WASH) services. Unsafe
water from all sources contributes significantly to the global burden of disease, principally through
the waterborne transmission of gastrointestinal infections such as cholera, typhoid, hepatitis, and a
wide range of agents that cause diarrhea and even death. Thus, cheap and effective water treatment
systems that can be used at different scales, from single-point water sources to small-community
water supplies, can make a valuable contribution to reducing the burden of disease by improving
access to safe water (Ahmed et al., 2011).
Microbiological contamination is a widespread problem and water is one of the most important
vehicles for disseminating this type of pollution, contributing to the dispersion of bacteria, yeasts,
fungi, spores, etc. Part of this contamination is the product of an uncontrolled discharge of
biological wastes or the usage of domestic sewage systems without the corresponding treatment.
Typically, these problems are very often solved with chlorine (or its derivatives) disinfec-
tion, an old, low cost water treatment technology that is very efficient and has an extensive use.
Alongside these advantages, it is well-known the existence of an important drawback resulting
from the toxicity of some of the chlorine disinfection by-products (DBPs) produced by the inter-
action of chlorine and chlorine derivatives with organic substances either naturally existing in
water, or resulting from improperly treated industrial or sanitary wastes (McDonnel and Russell,
1999). Some of these DBPs have been already included in the existing lists of substances having
mutagenic or carcinogenic properties.
During the last years, organizations of different origin have insisted in the need for a gradual
substitution of chlorine for water disinfection and requested for more research efforts aimed at
developing efficient alternatives having reasonable costs (Ahmed et al., 2010). Global reduction
of chemical deposition into the environment is necessary.
Addressing these problems calls out for a tremendous amount of research to be conducted to
identify robust new methods of purifying water at lower cost and with less energy, while at the
same time minimizing the use of chemicals and impact on the environment.
In the latest advances in water purification and disinfection, mainly in the oxidation of toxic
organic compounds, persistent and cumulative, are used the new technologies of advanced oxida-
tion processes (AOPs), which are methods that involved chemical or photochemical generation
and use of species transitional powerful as the hydroxyl radical (HO• ).
This work contains a comprehensive, albeit reduced, report on some of the processes in use,
the kinetic modeling that accompany several of them and the theories behind those proposals,
especially when they have been developed by us, in relation to technologies for water disinfection.
253
254 M. Flores et al.

Five disinfection methods were compared: (i) UV disinfection, (ii) hydrogen peroxide disin-
fection, (iii) peracetic acid disinfection (iv) peracetic acid + UV disinfection and (v) Hydrogen
peroxide + UV disinfection. The main target of the study was trying to understand and interpret
the differences that exist between the different procedures. In addition, we were searching for
quantitative information in order to get an idea, as approximate as possible, about operating con-
ditions and final results, with the aim of being able to distinguish among them, which might be
the most efficient, economical and environmentally friendly method.

15.2 DISINFECTION

Disinfection is the process used to reduce the number of pathogenic microbes in the water (US
EPA, 2003).
The most common and widespread health risk associated with drinking water is contamination,
either directly or indirectly, by human or animal excreta and the microorganisms contained in
feces. Drinking such contaminated water or using it in food preparation may cause new cases
of infection. Pathogenic (disease-causing) organisms of concern include bacteria, viruses and
protozoa.
The disinfection process has been routinely carried out since the dawn of the 20th century to
eradicate and inactivate the pathogens from water used for drinking purpose. Disinfectants in
addition to removing pathogens from drinking water, serve as oxidants in water treatment. They
are also used for (i) removing taste and color; (ii) oxidizing iron and manganese; (iii) improving
coagulation and filtration efficiency; (iv) preventing algal growth in sedimentation basins and
filters, and (v) preventing biological regrowth in the water distribution system (US EPA, 1989).
Disinfection may be chemical, physical or a combination of both. Many disinfectants are
used alone or in combinations (e.g. hydrogen peroxide and peracetic acid) in the health-care
setting. These include alcohols, chlorine and chlorine compounds, formaldehyde, glutaraldehyde,
ortho-phthalaldehyde, hydrogen peroxide, iodophors, peracetic acid, phenolics, and quaternary
ammonium compounds. A viable alternative could be the use of chemical agents plus UV radiation
to avoid revival of microorganisms. Disinfection kinetic models are the basis for assessing the
disinfection performance and the design of contactor systems (Trussell and Chao, 1977). Over
the years, a number of kinetic models have been proposed for the formulation of disinfection
design criteria.
Model adequacy is dependent upon the robustness of the underlying inactivation rate law. If
the model accounts for the disappearance, the most cited are: Chick (1908); Watson (1909); Gard
(1957); Selleck et al. (1978); Hom (1972), Hass (1999); Severin et al. (1983) among others.
A summary of some water disinfection processes with kinetic modeling and used devices that
have been developed by us (case studies) in Table 15.1.
All these experiments were conducted with pure water. In practical cases, usually the water
will have impurities. Both organic residues as well as inorganic salts affect the efficiency of the
process, either because they consume the oxidizing agent (when applicable) or because according
to their size, they can help bacteria to be concealed behind these substances affecting the capacity
of penetration of radiation when this treatment is applied. In any practical application, these
considerations will have to be born in mind.

15.3 UV DISINFECTION

UV disinfection is using the ultraviolet light with appropriate wavelengths to penetrate the cells of
microorganisms and destroying the molecular structure of DNA (deoxyribonucleic acid) or RNA
(ribonucleic acid). It results in growing cell death and (or) regenerative cell death. Thereby, the
microorganism cannot reproduce (Smith, 2011). UV disinfection is a physical method. It does not
add any substance to the water, and operates without any side effects. It is better than chlorination
Water disinfection with UVC and/or chemical inactivation 255

Table 15.1. Summary of cases in study.

Disinfection Microorganism
process model Reactor Kinetic model/mechanism approach

UVC Escherichia coli Well stirrer batch Modified Severin model (proposed by Labas
ATCC 8739 annular reactor(*) et al., 2009).
Hydrogen Escherichia coli Well stirrer batch Series parallel pseudo-chemical steps (proposed
peroxide ATCC 8739 annular reactor(*) by Flores et al., 2012).
Peracetic acid Escherichia coli Well stirrer batch An extended series parallel pseudo-chemical
ATCC 8739 annular reactor(*) steps (described in this work).
UV/Hydrogen Escherichia coli Well stirrer batch Modified Severin model (proposed by Labas
peroxide ATCC 8739 annular reactor(*) et al., 2009).
UV/Peracetic Escherichia coli Well stirrer batch Mechanistic approach (described in this work).
acid ATCC 8739 annular reactor(*)

(*) In all experiments, a well-stirred batch annular reactor having a total reaction volume of 2000 cm3 was
employed. The internal radius is ri = 3.7 cm and the external one is ro = 7.5 cm. Stirring was achieved
with a custom made, strong, eccentrically operated, orbital shaking device. A cooling jacket connected to
a thermostatic bath (Haake) surrounds the reactor to keep the reacting system at a constant temperature of
20◦ C. See also Figure 15.1.

Figure 15.1. Experimental reactor. The lamp shown in the figure was used only in those experiments that
required UVC radiation.

about in that aspect. It is usually combined with other substances, for example: UV + H2 O2 ,
UV + H2 O2 + O3 , UV + TiO2 , in order to obtain better disinfection results (US EPA, 2006).
Among the advantages of the use of UV disinfection of water can be mentioned:
• No addition of chemicals.
• Neither pH nor temperature-dependent.
• Specific inactivation mechanism.
• Effective against parasites

15.3.1 The principle of UV disinfection


According to Bolton and Cotton (2008), the mechanism of UV disinfection depends on the
absorption of radiation in proteins and nucleic acids (RNA and DNA) of a given microorganism.
256 M. Flores et al.

Figure 15.2. DNA damage UV-induced. (a) thymine-thymine cyclobutane pyrimide dimer and (b) thymine-
cytosine dimer. Where T: thymine and C: cytosine.

The absorption of large UV doses of the proteins present in the cell membranes leads to rupture
of these membranes and, consequently, to cell death. In contrast, absorption of lower UV doses of
the DNA may disrupt the ability of the microorganism to reproduce, preventing ulterior infection
of the medium.
The effective UV wavelength range can be divided into four different bands: UVA (400–
315 nm), UVB (315–280 nm), UVC (280–200 nm) and vacuum ultraviolet light (200–100 nm).
UVB and UVC light are always used for the disinfection of drinking water treatment, since they
have higher germicidal properties (Harm, 1980). However, the lamp market has been moving
slowly toward sources of radiation usually knows as germicidal lamps with preponderant almost
monochromatic radiation in the UVC (λ = 253.7).
Different forms of DNA lesion have been found to result from UVC-induced damage: some
molecules present in the DNA, such as purines and pyrimidines, absorb strong ultraviolet radiation
(maximum at 254 nm) and undergo chemical changes, such as dimers and hydrates. The two
major UV-induced DNA lesions are cyclobutane-pyrimidine dimers, CPDs, (see also Fig. 15.2)
and Dewar valence isomers (Häder and Sinha, 2005). The dimerization has been considered the
main cause of mutagenic effects resulting from UV radiation.

15.3.1.1 Repair mechanisms


Many organisms have developed active and passive preventive measures to minimize or repair
phototoxic DNA damage and can be repaired following UV disinfection by light-dependent (pho-
toreactivation) and light-independent (dark repair) mechanisms found in many organisms, such
as fecal coliforms, which include Escherichia coli. The repair mechanisms are:
Photoreactivation: The photoreactivation process is catalyzed by enzymes known as photolyases,
which directly reverse DNA damage, and consequently recovers microbial viability (Weber,
2005).
Recombination repair: recombination repair fills the daughter strand gap by moving a comple-
mentary strand from homologous region of DNA to the site opposite the damage (Häder and
Sinha, 2005).
Water disinfection with UVC and/or chemical inactivation 257

Excision repair: Dark repair pathway function by replacing the damaged DNA with a new,
undamaged nucleotides based on the information on the complementary DNA strand (Britt,
1996; Hader and Sinha, 2005).

15.3.2 Case study: UV disinfection in clear water conditions


This section work presents a detailed kinetic study of the rate of removal of a model bacteria
(Escherichia coli) employing UV radiation (253.7 nm) in a laboratory reactor where all the sig-
nificant operating variables were carefully measured and controlled. A significant modification
of the Series-Event Model was used to interpret the experimental data. The developed model is
based on a rather complex dependence with respect to the Escherichia coli concentration and to
the radiation that is effectively absorbed by the bacteria which was precisely quantified. The con-
centration evolution was analyzed employing the plate count method with Petrifilm(TM) specific
plates.

15.3.2.1 Experimental procedure


Escherichia coli strain ATCC 8739 was used throughout this work. The purity of the strain was
verified by conventional methods (APHA, 1984; Marshall, 1992).
The culture was grown in a complex medium (Nutrient broth) that had beef extract as the main
component. Therefore, the broth composition was: tryptone: 10 g L−1 , beef extract 5 g L−1 and
NaCl: 5 g L−1 .
The working solution was prepared from a culture that are in the beginning of the stationary
phase and afterwards was brought to a 1/1000 dilution with physiological saline to simulate clear,
transparent waters. The specific absorption coefficients of the two culture media and E. coli
were measured in a UV-Vis Lamda 40 Perkin-Elmer Spectrophotometer at 253.7 nm. An atomic
spectroscopy analysis (Perkin-Elmer-5000 AAnalyst) detected traces of iron and copper ions in
the growing culture (Cu = 7.7 µg g−1 and Fe = 43 µg g−1 ).
Initial E. coli concentrations ranged from 104 to 106 CFU cm−3 (CFU = colony forming unit).
Notice that, in this case, the system was irradiated, in distinct experiments, with two tubular
lamps of different power, placed on the centerline of the annular space and separate from the
polluted water by a concentric quartz tube.

15.3.2.2 Experimental runs


The lamps were turned on, allowing 30 minutes for stabilizing their operation. The sample at t = 0
was taken at the same time that the lamp shutters were taken off. Afterwards, samples were taken
at different time intervals for several measurements.
The exact value of the initial concentration of bacteria was measured for each experiment.
Runs were duplicated and samples subjected to triplicate determinations. The initial pH was 7
and remained practically constant during all the runs. It is necessary to emphasize that none of
these experiments was done in the presence of any oxidizing agent as for example, hydrogen
peroxide, in order to make sure that only the effect of the ultraviolet radiation was observed.
The CFU counting method was made with a sterile peptone water solution. Dilutions of the
samples to obtain the optimum concentration for the CFU counting were used. Each sample
was subjected to the following measurements: absorbance a 253.7 nm and CFU counting using
specific Pretrifilm™ plates (3 M Microbiology Products) for E. coli and coliform bacteria. The
plates were incubated during 24 hours at 37◦ C.
After one experimental run was completed, the UV lamp was turned on for 30 minutes. After-
wards, the reactor was washed with distilled water. This operation was repeated three times. Then
the reactor remained kept with a mixture of distilled water and alcohol until the next experimental
run. Before starting a new experiment, the reactor was washed again twice with distilled water
to remove any residual alcohol. Before beginning with the usual protocol corresponding to the
characteristics of every experiment, the UVC radiation source was turned on again for a period
258 M. Flores et al.

of an hour with the reactor completely empty. This procedure was applied in the four different
cases described in this work.

15.3.2.3 Kinetic model


The developed model is a modification of the Series-Event one originally proposed by Severin
(1982). The model is thought of as a series of consecutive “damaging reactions” each one of them
producing a partial alteration on the structure of the different chemical blocks that construct the
DNA and RNA sequence of the bacteria. However, it exist a threshold limit. Organisms which
have reached an event level greater than the threshold limit are inactivated and those which are
below this level survive. For this reason, part of the information needed for the model is to know
this threshold limit.
The original kinetic model was formulated in terms of mixed second order kinetics with respect
to the local light intensity and the concentration of microorganisms existing in each event level.
However, in the experiments performed in a batch reactor, assumption was made that the light
intensity was uniform and constant and the hypothesis was incorporated in the kinetic model
(uniform intensity).
Experiments were also carried out in a completely mixed, annular, flow-through reactor. This
assumption was fully removed in our work.
After a damaging event, it is considered that a new species is formed. Then, the process can be
presented as a series of n reactions according to (Labas et al., 2009):
k1 k2 ki+1 kn
CEc,0 −−−→ CEc,1 −−−→ CEc,i −−−→ CEc,n−1 −−−→ CEc,n (15.1)

with i = 0, 1, 2, . . ., i, . . ., n and k1 = k2 = kn = k
0
It should be noticed that the initial concentration of bacteria is CEc whereas CEc,0 is the concen-
tration of those that have not received any damage. Only at t = 0, CEc 0
= CEc,0 . At any other time
0
CEc,i < CEc . CEc,i is the concentration of a given state of damage (state i) in the microorganism
that has reached the level of injured “i” “and “n” is the threshold limit” equal to the number of
events needed to reach inactivation thus n represents inactivated “species” that have suffered an
usually quasi-irreversible damage an can be counted as dead bacteria. The total number of existing
“species” is n + 1 in order to account for the species that have not yet received any injury.
All bacteria, either the ones that have no damage or the others that have received some level
of damage, but have not reached the threshold limit “n”, are survival bacteria CEc. Because the
number of inactivated or death bacteria can be obtained from the difference between the initial
0
concentration CEc and the concentration of survival bacteria CEc

n−1
CEc,dead = CEc
0
− CEc = CEc
0
− CEc,i (15.2)
i=0

It is possible to write the equivalent to a mass balance (bacteria inventory) for each species:
For i = 0:
dCEc,i
= −ki+1 (CEc,i )δ (αλ Eλ,0 )γ  (15.3)
dt
For i = 1, 2, . . . , n − 1:
dCEc,i  
= ki (CEc,i−1 )δ (αλ Eλ,o )γ  − ki+1 (CEc,i )δ (αλ Eλ,o )γ (15.4)
dt
For i = n:
dCEc,i  
= ki (CEc,i−1 )δ (αλ Eλ,o )γ (15.5)
dt
Equations (15.3) to (15.5) need the value of the fluence rate as a function of position and time.
From the definition of the fluence rate:

Eλ,o (x, t) = Lλ, (x, t) d (15.6)

Water disinfection with UVC and/or chemical inactivation 259

Thus, it can be obtained if the value of the radiance (radiation power) is known. From the
radiative transfer equation for homogeneous media:
dLλ, (s,t)
+ αλ (s,t)Lλ, (s,t) = 0 (15.7)
ds
where Lλ, is the spectral radiance of wavelength λ and direction .
With the boundary condition:
Lλ, (sR ,t) = L0λ, (t) (15.8)
Equation (15.7) can be formally integrated and substituted into Equation (15.6). For conve-
nience, the solid angle may be written in terms of a spherical coordinate system, resulting in the
following result:
 φ2  θ2   sL 
Eλ,o (s, t) = dφ dθ sin θ (L0λ (θ, φ, t)) exp − α(s, t) ds (15.9)
φ1 θ1 sR

To calculate the radiation field inside the reactor, an emission model for the lamp is needed.
With this purpose, the three dimensional source with volumetric emission model proposed by
Cassano et al. (1995) will be applied. As shown in the Appendix for integrating Equation (15.9)
the following boundary conditions apply:
For the value of: L0 (θ, φ)
  0.5
Pλ,L 2[r 2 (cos2 φ − 1) + rL2 ]
L (θ,φ) = ϒw
0
(15.10)
4π 2 rL2 LL sin θ

ϒ w is an average value of the reactor wall transmittance. (Notice that the subscript λ has been
dropped because this work is carried out with monochromatic radiation).
And for the limiting angles are:
 
2 ½
−1 r cos φ − [r (cos φ − 1) + rL ]
2 2
θ1 (φ) = tan (15.11)
(LL − z)
 ½

r cos φ − [r2 (cos2 φ − 1) + rL2 ]
−1
θ2 (φ) = tan (15.12)
−z
 
−1 (r 2 − rL2 )½
−φ1 = φ2 = cos (15.13)
r
The final equation that can be solved with the help of numerical integration is:
⎧φ ⎡ s ⎤⎫
2 θ2 L
(ϒw Pλ,L )(4π 2 rL2 LL )sin θ ⎨ 2 0.5 ⎣ ⎦

Eo (r, z, t) = (dφ) dθ 2[r 2
(cos 2
φ − 1) + r ] exp − α(s, t)ds
2[r 2 (cos2 φ − 1) + r 2 ] ⎩ ⎭
0.5 L
L φ1 θ1 sR

(15.14)
However, the value of Pλ,L is not always well known (it is provided by the lamp manufacturer)
and invariably changes along the time of operation of the useful life of the radiation source. When
the fluence rate at the reactor wall can be measured with actinometer methods (Murov et al.,
1993; Zalazar et al., 2005), it is convenient to use an alternative form of Equation (15.14). This
is particularly always possible in laboratory research work.
 
lamp reactor characteristics
E0|W  = f (15.15)
System geometry
with E0|W  being the fluence rate at the reactor wall.
260 M. Flores et al.

Therefore, with the help of the elaborate algebraic artifice it is possible to transform Equation
(15.14) into the following expression:

⎡ ⎤
φ2 θ2 sL
 Eo |W 
dθ 2[r 2 (cos2 φ − 1) + rL2 ] exp⎣− α(s,t)ds⎦
0.5
Eo (r, z, t) = dφ (15.16)
ψ
φ1 θ1 sR

where ψ is a geometric factor that always should be computed because LR , rL and ri are known:
LR φ2 (ri ) θ2 
(φ,ri ,z)
1 0.5
ψ= dz dφ dθ 2[r 2 (cos2 φ − 1) + rL2 ] (15.17)
LR
0 φ1 (ri ) θ1 (φ,ri ,z)

Notice the difference between Equations (15.14) and (15.16); in the second, ri is a constant.
Eo |W /ψ gives the boundary condition just before the absorption process produced by the reacting
medium commences at r = ri . ψ accounts for the geometrical relationship that expresses the
relative location of the lamp (with all its dimensions) and the inner wall of the reactor at r = ri .
It is clear that, with the exception of the wall compound transmission coefficient given by ϒW ,
from the lamp until the point at r = ri , the medium is transparent. Eo |W is the value that can be
obtained with potassium ferrioxalate actinometry (Murov et al., 1993; Zalazar et al., 2005).
Finally, since for a reactor of constant cross sectional area, the volume average is reduced to an
integral over the reactor length, the volume average of Eo (r, z, t) must be calculated according to:

1 LR
Eo (r,z,t) = Eo (r,z,t) dz (15.18)
LR 0

The final result to be used in Equations (15.3) to (15.5) is:


⎧ ⎡ s ⎤⎫γ
 γ LR ro ⎨φ2 θ2 L ⎬
  2 Eo |W
dφ dθ 2[r 2 (cos2 φ − 1) + rL2 ] exp⎣− α(s, t)ds⎦
0.5
Eo = 2
γ
dz r dr
(ro − ri2 )LR ψ ⎩ ⎭
0 ri φ1 θ1 sR

(15.19)

The linear napierian radiation absorption coefficient of the system (α) is defined as:

α = αEc + αc with αEc = κEc,i CEc,i (15.20)

 λmax
1
and αc = αλ Pλ dλ (15.21)
Pc λmin


where κ[=]cm2 (CFU)−1 and CEc,i [=]CFU cm−3 .
In order to facilitated the understanding of the development of the limits of integration of
Equation (15.19), in the appendix included at the end of the main text, it has been added a figure
to explain clearly the meaning of all the variables used to obtain Equations (15.10) to (15.13).

15.3.2.4 Experimental results


Disinfection with UV radiation has been effective, achieving high rates of inactivation at short
contact times (Fig. 15.3) employed worldwide for decades. The primary advantage of UV disinfec-
tion is to control microorganisms without chemicals and it has numerous benefits for municipal,
industrial and commercial customers.
Water disinfection with UVC and/or chemical inactivation 261

Figure 15.3. Results employing UVC radiation, with different lamps. Dotted lines and solid lines are results
from the model. (Reproduced from Photochemical & Photobiological Sciences; Reference:
Labas et al., 2009. Reproduced by permission of The Royal Society of Chemistry: www.
res.org/pbs).

15.4 HYDROGEN PEROXIDE

The antimicrobial and/or antiseptic properties of hydrogen peroxide have been known for many
years because of its efficacy and reasonable manipulation safety (Dalrymple et al., 2010; Labas
et al., 2006). It is effective against a wide spectrum of bacteria, yeast, molds, viruses and spore
forming organism (Cords et al., 2005; Labas et al., 2008). The cytotoxic effect exerted by hydrogen
peroxide on microorganisms depends on the cell type used, its physiological state, length of
exposure, environmental condition, H2 O2 concentration used and the cell culture media employed
(Labas et al., 2008; Raffellini et al., 2010). According to its concentration hydrogen peroxide
can act as bacteriostatic or as bactericide. The cytotoxic effects of H2 O2 on Escherichia coli have
been extensively studied (Labas et al., 2005; 2008; Raffellini et al., 2010).
For water disinfection purposes, the non-persistent characteristic of hydrogen peroxide becomes
a disadvantage to maintain the water quality in the distribution system. However, its use is widely
spread because it is relatively inexpensive, is easily removed when desired and is unlikely to be
health hazardous if used properly.

15.4.1 The principle of disinfection using hydrogen peroxide


Hydrogen peroxide is disinfectant with recognized antimicrobial properties that can be used
because is effective and easy to manipulate. In addition, the chemical mechanisms that promote the
hydrogen peroxide decomposition can be inferred but infrequently established with indisputable
certainty. However it is known that is due to its ability to generate strongly oxidant chemical species
like the hydroxyl radical HO• . This radical species reacts with almost all biological molecules.
The attack by the HO• radical, in the presence of oxygen, initiates a complex cascade of oxidative
reactions leading to decomposition of the all the enzymes and organic compounds that result from
the rupture of the cell membrane.

15.4.2 Case study: hydrogen peroxide disinfection in clear water conditions


This section is an attempt to model the disinfection process based on a chemical interpretation
of the possible mechanism of cell damage. The key point is to search for the most accepted
assumptions concerning the possible place where the action of reactive hydroxyl radical can
take place. According to literature, (Dalymple, 2010; Maillar, 2002; Mc Donnell and Russell,
1999) in the case of bacteria three sites of the cell could be the targets for HO• invasion: (i)
262 M. Flores et al.

The peptidoglycan layer, (ii) the lipopolysaccharide layer (found only in Gram-negative bacteria)
and (iii) the phospholipid bilayer. In this study Escherichia coli was the chose bacteria and
consequently the three layers will be present.
After the attack to the membrane, the oxidation of the products resulting from the lysis of the
bacteria, was modeled as a series of chemical events, which leads to its dead or to an irreversible
damage. The pathway of the kinetic model includes: active (BAC ), inactive (BIN ) and death (BDE )
population of bacteria, as well as several additional chemical products of the lysis with the generic
denomination of LYP1 , LYP2 , . . . , LYPn .
The kinetic model developed in this work was successfully validated with experimental data.

15.4.2.1 Experimental procedure


The experimental procedure, the culture media preparation and handling and conditioning of the
bacterium has been detailed in 15.3.2.1.
The following chemical were used in this experience: hydrogen peroxide (Merck, pro analysis
30%), Catalase (from Aspergillus niger, Biochemika), Ammonium molybdate (Cicarelli, pro-
analysis), Potassium iodide (Cicarelli, pro-analysis), Potassium acid phthalate (Cicarelli, 95%
pro-analysis), Physiological saline solution (Roux-Ocefa), Nutrient broth (Biokar Diagnostics),
Sodium hydroxide (Cicarelli, pro-analysis) and Peptone water (Biokar Diagnostic).
To quench the hydrogen peroxide action during the time interval between sampling and spread
plating, a known fraction of the sample was mixed with the required amount of catalase solu-
tion. Control experiments were conducted to ensure that the employed concentrations of catalase
solutions did not affect bacterial concentrations. After spreading the plates with the appropriate
volume of sample they were incubated for 24 h at 37◦ C.

15.4.2.2 Kinetic model


This development is an attempt to model the disinfection process based on a chemical
interpretation of the possible mechanism of cell damage.
In modeling the intrinsic reaction kinetics of disinfection with hydrogen peroxide one of the
major difficulties is to include the iron or iron-superoxide participation in the mechanism to
produce hydroxyl radicals (Kehrer, 2000; Sies, 1991); i.e. the Fenton-like or Haber-Weis-like
reaction:
k1
H2 O2 −−−−− −−−−−−−− −−−

−−−−−−→ OH− + HO• (15.22)
Fe +H2 O2 or Fe +O2 promotion
2+ 3+

Note that, when the reaction is written as above, without including specifically the Fe3+ or the
Fe + O•−
3+
2 promotion, k1 is not known.
The formation of hydroxyl radicals is proposed according to the following path:
Propagation:
k2
H2 O2 + HO• −−−→ HO•2 + H2 O (15.23)
k3
H2 O2 + HO•2 −−−→ HO• + H2 O + O2 (15.24)
Termination:
k4
2HO• −−−→ H2 O2 (15.25)
k5
2HO•2 −−−→ H2 O2 + O2 (15.26)

k6
HO• + HO•2 −−−→ H2 O + O2 (15.27)
15.4.2.2.1 Membrane disruption
The process that ends up with the membrane breakdown may be treated by resorting to a pseudo
homogeneous interpretation of the intricate network of superficial reactions that leads to the
rupture of the protective envelope. Let HSCW|B represent a hypothetical species of the components
of the cell wall of the active bacteria whose concentration can be expressed in units of mol cm−2
Water disinfection with UVC and/or chemical inactivation 263

and that reacts with the HO• radical. This composition is, on the average, the same one for every
specific bacterium species. Then,
k7∗
HSCW |BAC + HO• −−−→ HSCW |BIN (15.28)

At this point, it is possible to think of SSA|BAC as the specific superficial area per unit volume of
one cell (in units of cm2 cm−3 ), V|BAC as the volume of one cell per CFU (in units of cm3 CFU−1 )
and [BAC ] as the whole instantaneous concentration of the active bacteria (in units of CFU cm−3
of reacting medium). The concentration of hypothetical species per unit volume of the fluid may
be calculated from the abovementioned superficial concentration of this species as:
[ HSCW |B ] × SSA|B × V |BAC × [B ](t) (15.29)
) *+ AC, ) *+ AC, ) *+ , ) AC
*+ ,
Superficial concentration of Superficial area per unit volume of Instantaneous concentration of
hypotetical species volume of bacterium one CFU active bateria per unit volume
of one bacterium
) *+ ,
Instantaneous representation of the total volumetric
concentration of hypotetical species of the active bacteria

Taking into account averaging values specific bacterium species, [HSCW |BAC ] in the cell wall,
SSA|BAC and V |BAC may be assumed almost constant. Then:
[HSCW |BAC ] × SSA|BAC × V |BAC = constant (15.30)
- .
k7 = k7∗ × HSCW |BAC × SSA|BAC × V |BAC (15.31)
)*+, )*+,
Pseudo homogeneous Superficial
volumetric kinetic constant kinetic constant
However, it is very unlikely that just one hydroxyl radical will produce a serious damage to
the cell wall. Neither the number of moles of which HO• that are necessary to consider that
the bacterium has been injured, nor the number of bacterium that form a CFU are known. It is
always possible to conceive an oxidation yield as the ratio of injured CFU with respect to the
spent hydroxyl radicals to produce this event:
k7 Injured CFU CFU cm3
k7# = ; Y7 = [=] ; k7# [=] (15.32)
Y7 HO• spent in this event mol CFU s

In order to have the proper application of this yield as well as achieving unit’s homogeneity it
must be considered that:
R7,OH = −k7∗ [BAC ][HO• ] (15.33)
With this pseudo homogeneous, biological reaction approach, step 15.28 can be finally
expressed as:
k7
BAC + HO• −−−→ BIN (15.34)
In a similar way, the same procedure can be applied to the injured bacteria in the reaction with
the HO• radicals according to:
- .
k8 = k8∗ × HSCW |BIN × SSA|BIN × V |BIN (15.35)
)*+, )*+,
Pseudo homogeneous Superficial
volumetric kinetic constant kinetic constant

k8 Dead CFU CFU cm3


k8# = ; Y8 = • [=] ; k8# [=] (15.36)
Y8 HO spent in this event mol CFU s
R8,OH = −k8# [BIN ][HO• ] (15.37)
k8
BIN + OH −−−→ BDE
g
(15.38)
This step leads to the rupture of the cellular wall and the death of the bacterium.
264 M. Flores et al.

Lysis of dead bacteria


The death of the bacterium implies that the cellular wall has been broken permitting the access
of the oxidant to the intracellular components, rendering the lysate:
k9
BDE + HO• −−−→ LYSATE (15.39)

Once the anatomic morphology of the cell has been altered producing the lysate, the HO•
can interact with its internal components in typical homogeneous reactions. From the chemical
point of view the activity of the HO• radicals can be interpreted in terms of several parallel-series
reactions with an undefined number of hypothetical components of the resulting lysis. In the
model they are represented by LYP,1 , LYP,2 ,…,LYP,n−1 , LYP,n that correspond to important groups
of compounds which, after ulterior oxidations will produce the end products of the disinfection
process:
+HO• +HO• +HO•
LYP,1 −−−−−−→ LYP,1 −−−−−−→ · · · · · LYP,1,n−1 −−−−−−→ EP1,n (15.40)
• • •
+HO +HO +HO
LYP,n−2 −−−−−−→ LYP,n−2,1 −−−−−−→ · · · · · LYP,n−2,n−1 −−−−−−→ EP,n−2,n (15.41)
• • •
+HO +HO +HO
LYP,n - 1 −−−−−−→ LYP,n−1,1 −−−−−−→ · · · · · LYP,n−1,n−1 −−−−−−→ EP,n−1,n (15.42)
• • •
+HO +HO +HO
LYP,n −−−−−−→ LYP,n,1 −−−−−−→ · · · · · LYP,n,n−1 −−−−−−→ EP,n,n (15.43)

15.4.2.3 Mathematical model final equations


The final expression for the disappearance rate of active bacteria is:
γA [BAC ][H2 O2 ]
RBAC = − (15.44)
[H2 O2 ] + γ0 [BAC ] + γ1 [BIN ] + γ2 [BDE ]

where:  
k 1 k7 k7# k8# (k9# + αk10 )
γA = ; γ0 = ; γ1 = ; γ2 = (15.45)
k2 k2 k2 k2
The final expression for the disappearance rate of injured bacteria is:
[BAC ][H2 O2 ]γA − [BIN ][H2 O2 ]γIN
RBIN = (15.46)
[H2 O2 ] + γ0 [BAC ] + γ1 [BIN ] + γ2 [BDE ]

where:  
k 1 k8
γIN = (15.47)
k2

15.4.2.4 Experimental results


According to Figure 15.4, it can be noticed that the model reproduces the experimental results.
The model also provides good information about the way in which the inactivation reaction
proceeds. It is possible to follow the changes in concentration of the active, injured and dead
bacteria, it is shown in Figures 15.4a,b for two different concentrations of hydrogen peroxide. In
the present context, it appears capable of accommodating biological recovery, as manifested by
the shoulder which appears in many survival curves.
From these two figures, it can also be observed a consistent behavior. When lower concen-
trations of hydrogen peroxide are used (45 ppm) the survival of a major number of active and
injured bacteria is seen, indicating that the oxidation is not sufficiently intense to overcome the
resistance of the cell wall. Conversely, at higher concentrations, (185 ppm) the number of survival
and undamaged bacteria gradually decreases and there is a breaking of the bacteria envelope in a
shorter time. Then, very rapidly, it changes from the state of injured bacteria to one of irreversible
death.
Water disinfection with UVC and/or chemical inactivation 265

Figure 15.4. Graphical description of the concentration evolution of the principal species existing in the
reacting medium as seen by the model. The plot follows changes in concentration of the
active, injured and dead bacteria at (a) 45 µg/cm3 and (b) 185 µg/cm3 of concentration of
H2 O2 . Experimental data: Solid line: culturable bacteria (BAC + BIN ); broken line: Active
bacteria; broken and dotted line: Injured bacteria; dotted line: Dead bacteria. (Reproduced
from Chemical Engineering Journal, Vol. 198–199/edition N◦ 1, Flores et al.; Pages 388–396;
Reproduced with permission from Elsevier).

15.5 PERACETIC ACID

Peracetic acid (PAA) use increased in the recent years because of its ecologically beneficial
properties (the reaction products are oxygen, water, and acetic acid) and its relative low cost.
However, the most remarkable attributes of PAA are: the broad spectrum of activity even in the
presence of heterogeneous organic matter, the absence of persistent toxic or mutagenic residual
266 M. Flores et al.

byproducts, no quenching requisites, small dependence of pH, short contact time requirements
and effectiveness for primary and secondary effluents. PAA is also characterized by its easy
technical preparation. The equilibrium state of its commercial solution is shown in the following
equation:
H+
−−
CH3 COOOH + H2 O ←−−−−
−−−−

−− CH3 COOH + H2 O (15.48)
PAA is a strong oxidant and disinfectant. Its oxidation potential is larger than that of chlorine or
chlorine dioxide. PAA is a more potent antimicrobial agent than hydrogen peroxide, being rapidly
active at low concentrations against a wide spectrum of microorganisms (Baldry, 1983; Baldry and
French, 1989b; Fraser et al., 1984). Its demonstrated effectiveness against V. cholera suggested
it should be a significant element in cholera control efforts. It is know that the PAA disinfection
capabilities are due to its ability to generate strongly oxidant chemical species such as the O− 2
superoxide radical or its conjugated base HO•2 , and the hydroxyl radical HO• . The damaging
effects of the bacteria cellular components seems to be produced by a particular phenomenon
called oxidative stress, resulting from those reactive oxygen species known as ROS (Labas et al.,
2006).

15.5.1 PAA mode of action


It is suggested that PAA disrupts the chemiosmotic function of the lipoprotein cytoplasmic mem-
brane and transport through dislocation or rupture of cell walls (Baldry and Fraser, 1988; Leaper,
1984). Thus, it may be that it is equally effective against outer membrane lipoproteins, facili-
tating its action against Gram negative cells (Leaper, 1984). Its action as a protein denaturant
may help to explain its characteristics as a sporicide and ovicide (Block, 1991). Furthermore,
intracellular PAA may also oxidize essential enzymes; thus, vital biochemical pathways, active
transport across membranes, and intracellular solute levels are impaired (Fraser et al., 1984).
It was demonstrated that PAA acts on the bases of the DNA molecule (Tutumi et al., 1973). An
important advantage of PAA is that it may inactivate catalase, an enzyme known to hinter free
hydroxyl radicals (Block, 1991).

15.5.2 Case study: water disinfection with peracetic acid in clear water conditions
This study was aimed at evaluating the disinfection efficiency of the PAA commercial solution
(15%) with the usual indicator of fecal contamination, Escherichia coli. The disinfection aptitude
of PAA was studied at different concentrations (1, 1.5, 2, 3, 4, 5, 6 mg L−1 ) as well as various
inactivation times. The reacting system used in all experiments was an annular, well-mixed batch
reactor having a total volume of 2000 cm3 . The feasibility of PAA for water efficient water
disinfection has been verified in this work: a 99.99% reduction of E. coli CFU was achieved,
with doses ranging from 1 to 6 mg L−1 and 5 minutes of contact time. A kinetic disinfection
mechanism is proposed to explain the obtained results.

15.5.2.1 Experimental procedure


The experimental procedure, the culture media preparation and handling and conditioning of the
bacterium has been detailed in 15.3.2.1.
The following chemical were used: Peracetic acid (Quimica Agroindustrial Neo: PAA 15%
v/v; H2 O2 20% v/v; acetic acid 25% v/v and water 40% v/v); Catalase (from Aspergillum
niger, Biochemika); Potassium permanganate solution, 0.1 N (Cicarelli, pro-analysis);
Sulfuric acid (Cicarelli, pro-analysis); Sodium thiosulfate (Cicarelli, pro-analysis); Physiological
saline solution (Roux-Ocefa); N,N-diethyl-p-phenylene-diamine (DPD) (Hach), Nutrient broth
(Biokar Diagnostics), Eosine Methylene Blue (EMB) Agar (Biokar Diagnostics) and Peptone
water (Biokar Diagnostic).
To quench the peracetic acid and hydrogen peroxide action during the time interval between
sampling and spread plating, a known fraction of the sample was mixed with the required amount
Water disinfection with UVC and/or chemical inactivation 267

of sodium thiosulfate and catalase solution respectively. Control experiments were conducted
to ensure that the employed concentrations of thiosulfate and catalase solutions did not affect
bacteria concentrations. The plates were incubated, after spreading them with the appropriate
volume of sample, for 24 h at 37◦ C in EMB plate.

15.5.2.2 A proposed kinetics of peracetic acid decomposition


Rokhina et al. (2010) developed a complete chain reaction mechanism that originated several
radical species, which can oxidize different sites (targets) of the bacteria’s cell membrane and its
internal constituents. The reaction was applied to advanced oxidation processes (AOPs), such as
phenol degradation with the assistance of MnO2 , and the cycle was analyzed both theoretically
and experimentally. The proposed scheme also resorted to previous reports that reinforced their
proposal Heywood et al. (1961), El-Agamey et al. (2003), Shi and Li (2007) and Ciotti et al.
(2008). Leaving aside the details that can be found in the original publication, the chain reaction
pathway is described as follows:

CH3 C(=O)OOH → CH3 C(=O)O• + HO• (15.49)


• •
CH3 C(=O)OOH + HO → CH3 C(=O) + O2 + H2 O (15.50)
• •
CH3 C(=O)OOH + HO → CH3 C(=O)OO + H2 O (15.51)

CH3 C(=O)O• → H3 C• + CO2 (15.52a)


• •
2CH3 C(=O)O → 2H3 C + 2CO + O2 (15.52b)

H3 C• + O2 → OOCH3• (15.53)
• •
CH3 C(=O)O + HO → CH3 C(=O)OOH (15.54)

Equation (15.53) is only important in oxygen saturated environments. It was found that Equation
(15.49) is the rate controlling step. The authors claim that all the generated radical species are
active contributors to the degradation mechanism but HO• and to some extent the H3 C radicals,
are the most significant ones. The reaction requires the presence of an eligible catalyst that should
be of the types usually encountered in Fenton or Fenton-like reactions. In our reacting system
employingAAS (Perkin-Elmer-5000AAnalyst) we always found traces of cooper and iron species.
Thus, this reaction could provide the necessary conditions for a fast attack to different components
of the microbial cellular membrane. Additionally, at pH between 5 and 10 the following reaction
is also possible (Koubet, 1964; Yuan et al., 1997):

CH3 C(=O)OOH + CH3 C(=O)OO− → 2CH3 C(=O)O• + O2 + H+ (15.55)

Equation (15.55) contributes to increase the concentration of oxidative radicals.

15.5.2.3 Experimental results


In Figure 15.5, it is shown that for concentrations larger than 5 ppm 99.9% disinfection is obtained
in less than 2 minutes. It is interesting to note that just small concentrations of PAA are needed
to produce these results.
However, it must be noted that these results were obtained using clear water and in the absence
of natural organic matter (NOM). PAA has been shown to be a very active oxidant of many
organic compounds and, consequently, the presence of this type of impurities in natural waters
will certainly have a negative influence in the actual inactivation effectiveness. In fact, they will
participate in parallel oxidative pathways competing with the attack produced by the oxidizing
species on active and/or partially injured bacteria. Other adverse effects such as for example, the
occurrence of inorganic salts, will also affect the disinfection mechanism.
268 M. Flores et al.

Figure 15.5. Disinfection results as a function of time. PAA is the parameter.

15.6 PERACETIC ACID + UV LIGHT

In the literature, the principal applications of these processes refer to the oxidation of organic
compounds, dissolved inorganic compounds and other pollutants that are toxic and/or refractory
to biological treatments. Few references were found, however, regarding their use for wastewater
disinfection (Caretti and Lubello, 2003).
Contrary to other advanced oxidation processes (UV/H2 O2 , O3 /H2 O2 , O3 /UV, TiO2 /UV),
numerous bibliographic references for the combined treatment between peracetic acid (PAA)
and UV do not exist (Caretti and Lubello, 2003). PAA based product consist of equilibrium
mixtures of peracetic acid, hydrogen peroxide, acetic acid and water in different proportions
where hydrogen peroxide plays different roles, being implied in the restoration of the equilibrium
between the different species after consumption of PAA but also acting as an oxidizing biocide
itself (Bianchini et al., 2002).
The biocide action of PAA and H2 O2 can be attributed to the production of highly reactive
radicals, above all the hydroxyl radical HO• originated by the cleavage of the peroxidic bond.
In the absence of UV irradiation, Fenton reactions, which are catalyzed by transition metal ions
traces are usually responsible for radical creation. In the presence of UV irradiation, radicals can
be photochemically produced by the cleavage of the O-O bond by UV light (Bianchini et al.,
2002).

15.6.1 Case study: disinfection of water with peracetic acid and its combination with UVC
15.6.1.1 Experimental procedure
The experimental procedure, the culture media preparation and handling and conditioning of the
bacterium has been detailed in Section 15.3.2.1.
Samples were taken every 1 second, with a sampling device especially designed for that purpose.
En all the disinfection experiments the plates were incubated, after spreading them with the
appropriate volume of sample, for 24 h at 37◦ C in EMB plate.
As mentioned in 15.3.2.1, in this case, the system was irradiated, in distinct experiments, with
two tubular lamps of different power, placed on the centerline of the annular space and separate
from the polluted water by a concentric quartz tube. In these cases, the lamps were turned on,
Water disinfection with UVC and/or chemical inactivation 269

allowing 30 minutes for stabilizing their operation. The sample at t = 0 was taken at the same time
that the lamp shutters were taken off. Afterwards, samples were taken at different time intervals
for several measurements.

15.6.1.2 A proposed kinetics of peracetic acid + UV


There are different possible conceptually different explanations to interpret the result of disin-
fection of Escherichia coli with PAA and UV radiation. We will advance some possibilities and
finally propose our tentative explanation of our data.
According to Caretti and Lubello (2003), it is possible to go back to the peroxide acids theory to
show how the action of UV produces a homolytic rupture in the O - O bond of the PAA molecule,
with the subsequent formation of the hydroxyl radical:
+hν
CH3 CO3 H −−−→ CH3 CO•2 + HO• (15.56)

CH3 CO•2 rapidly declines forming CH•3 and CO2 while the molecule of peracetic acid can sub-
sequently react with the HO• radicals produced, according to the following reactions of addition
and subtraction of labile hydrogen:

CH3 CO3 H + HO• → CH3 CO4 H2 + HOO• + CH3 CO2 H (15.57)

CH3 CO3 H + HO• → CH3 CO• + O2 + H2 O (15.58)

The presence of hydrogen peroxide within the commercial product of the PAA contributes not
only to the formation of new PAA as soon as it is consumed, but also to the formation of new
hydroxyl radicals.
According to Keller et al. (2008), the photodissociation of the PAA can follow one or even
more than one of the following three possible reaction pathways:

CH3 CO3 H−→CH3 CO•2 + HO• → CH3 + CO2 + HO• (secondary) (15.59)

The first reaction (15.59) is initiated by a homolytic O-O bond cleavage to generate two
photoproducts.
Equation (15.60) follows a concerted mechanism with simultaneous fission of an O-O bond
and a C-C bond:
CH3 COH → CH3 + CO2 + HO• (15.60)
Equation (15.61) is also stepwise reaction but instead begins with a C-C bond cleavage:

CH3 CO3 H → CH3 + CO3 H → CH3 + CO2 + HO• (secondary) (15.61)

15.6.1.3 Experimental results


As shown in Figure 15.6, the combination of UV/PAA achieved higher levels of inactivation at
lower contact times.
Results of the microbiological analysis on samples showed that PAA and UV together were
very effective in bacteria removal. It can be assumed, in accordance with previous mechanistic
proposals, that this very high effectiveness was due to the generation of hydroxyl radicals by the
photolysis phenomena, but the obtained inactivation rates were similar to those produced using
UV radiation as a single agent (7–8 seconds). This would indicate that the addition of PAA does
not lead to a significant increase in the effectiveness of the process. Nevertheless, the benefits of
this combination are that peracetic acid would act as a bactericide, preventing bacterial regrowth,
turning the bacteria inactivation an irreversible dead.
270 M. Flores et al.

Figure 15.6. Results employing UVC plus PAA.

15.7 HYDROGEN PEROXIDE + UV

Publications concerning kinetic studies of the combined use of UVC radiation and hydrogen
peroxide are very limited. Sundstrom et al. (1992) proposed two kinetic models: a mixed second
order model and an application of the typical Series-event model. Later Gardner and Sharma
(1992) described the inactivation of spores of B. subtilis in terms of a kinetic expression derived
from the Multi-Target model. Just recently, Alkan et al. (2007) studied inactivation working with
coliform bacteria in superficial waters and interpreted their results calculating the coefficients of
the Chick–Watson model.
However, there is no consensus concerning the most certain mechanism that explains the
action of H2 O2 /UV in disinfection process. Sundstrom et al. (1992) employing bacteria such as
Escherichia coli and Bacillus subtilis found a beneficial effect on the rate because of the addition
of hydrogen peroxide on the other hand, Rajala-Mustonen and Heinomen-Tanski (1995) found
the opposite result.
Rincon and Pulgarin (2004) found a synergistic effect of low wavelength UV radiation and low
concentrations of hydrogen peroxide (less than 10 ppm). Bayliss and Waites (1979) and Standard
et al. (1983) found an acceptable effectiveness of the method to inactivate vegetative cells and
spores employing rather large concentrations of H2 O2 .

15.7.1 Case study: disinfection with hydrogen peroxide and UV light in


clear water conditions
15.7.1.1 Experimental procedure
The experimental procedure, the culture media preparation and handling and conditioning of the
bacterium has been detailed in 15.3.2.1 and 15.3.2.2.
The lamps were turned on, allowing for 30 minutes stabilizing their operation. The working
solution was added to the reactor, immediately after the proper dosage of peracetic acid is added to
the reactor and immediately starts stirring to homogenize the medium. Samples were taken every
1 second, with a sampling device designed for that purpose. En all the disinfection experiments
the plates were incubated, after spreading them with the appropriate volume of sample, for 24 h
at 37◦ C in EMB plate.
Water disinfection with UVC and/or chemical inactivation 271

Table 15.2. Processing time to reach 99.99% reduction


of the CFU concentration.
Process Time/s

UVC (40 W) irradiation 10


UVC (15 W) irradiation 14

This table was adapted and modified from Labas et al.,


2009. Photochem. Photobiol. Sci. rsc.org/pps.

Figure 15.7. Results employing UVC plus hydrogen peroxide. 15 W nominal input power lamp. Dot-
ted lines, dotted and broken lines, broken lines and solid lines are results from the model.
(Reproduced from Photochemical & Photobiological Sciences; Reference: Labas et al., 2009).
Reproduced by permission of The Royal Society of Chemistry: www. res.org/pbs).

15.7.1.2 Kinetic model


The Modified Series-event model was applied once more. The equations for this process are very
similar to those derived for the first case (Section 15.3).
However, Equation (15.16) in Section 15.3 must be significantly modified according to
(Labas et al., 2009):
α = αEc + αc + αp (15.62)
where αp = κp Cp . In these equations, the linear napierian radiation absorption coefficients are
defined for the bacteria, the aqueous solution and hydrogen peroxide respectively, each one of
them derived from the corresponding values of concentrations and κ expressed in terms of cm2
mole−1 or cm2 CFU−1 .

15.7.1.3 Experimental results


Table 15.2 shows the time required to reach the same level of reduction of the microbiological
contamination for three of the alternatives explored before. It is clear that the effect produced by
hydrogen peroxide alone is negligible compared with the one corresponding to UVC alone.
Then, under these conditions, it is not necessary to consider the existence of two separate
reaction kinetics that compete in parallel.
However, a very significant outcome was observed. The time required to reach 99.99% inacti-
vation increases when H2 O2 is added. The larger the added concentration of hydrogen peroxide,
the longer is the time to reach the desired inactivation results.
One explanation is immediate: hydrogen peroxide acts as an inner filter, absorbing UVC
radiation (at 254 nm, its molar napierian absorption coefficient is not too high but important as
compared with the other components of the reacting medium).
The combination of UVC and hydrogen peroxide, under the explored operating conditions, is
detrimental to the UVC effectiveness performance (Figs. 15.7 and 15.8).
272 M. Flores et al.

Figure 15.8. Results employing UVC plus hydrogen peroxide. 40 W nominal input power lamp. Dot-
ted lines, dotted and broken lines, broken lines and solid lines are results from the model.
(Reproduced from Photochemical & Photobiological Sciences; Reference: Labas et al., 2009.
Reproduced by permission of The Royal Society of Chemistry: www. res.org/pbs).

15.8 CONCLUSIONS

Disinfection by ultra-violet light (UV) is accepted as being the most effective in removing
pathogenic organisms. However, since the process typically consumes a great deal of power it is
both expensive to operate and has a large carbon footprint. Depending on the financial resources
existing, the water distribution rate needed and the quality of the source water the choice of
treatment may vary.
Throughout the industrialized world and advanced developing countries, large efforts are made
investigating alternative disinfection/oxidation practices. This chapter is a brief account of the
possibilities of peracetic acid and hydrogen peroxide as additional optional choices for water
treatment plants.
From this point of view, our preliminary results seem to suggest that peracetic acid may be
considered as a good alternative oxidant for water treatment processes of water, since potentially
produces less unwanted byproducts, becoming an environmentally friendly agent.
A second final consideration indicates that it is important to complete the development of very
rigorous and reliable modeling of all the involved kinetics. This means to carry out work with
waters having more realistic compositions; i.e., considering cases that may contain organic matter
or inorganic salts that could severely affect the process efficiency. This additional information
will facilitate the proposal of reactor design and scaling-up methods that are necessary to help
out with the application of commercial competitive processes.

ACKNOWLEDGEMENTS

The authors are grateful to Universidad Nacional del Litoral (UNL), Consejo Nacional de Inves-
tigaciones Científicas y Técnicas (CONICET), and Agencia Nacional de Promoción Científica
y Técnica (ANPCyT) for its financial support. We also thank the editorials that allowed us the
reproduction of some graph.

Graphics 15.4.1 (a) Reprinted from Chemical Engineering Journal , Vol. 198–199/edition N◦ 1,
Authors: Marina J. Flores, Rodolfo J. Brandi, Alberto E. Cassano, Marisol D. Labas, “Chemical
disinfection with H 2 O2 – The proposal of a reaction kinetic model”, Pages 388–396, Copyright
(2013), with permission from Elsevier.
Water disinfection with UVC and/or chemical inactivation 273

Graphics 15.4.1 (b) Reprinted from Chemical Engineering Journal , Vol. 198–199/edition N◦ 1,
Authors: Marina J. Flores, Rodolfo J. Brandi, Alberto E. Cassano, Marisol D. Labas, “Chemical
disinfection with H 2 O2 – The proposal of a reaction kinetic model”, Pages 388–396, Copyright
(2013), with permission from Elsevier.
Graphic 15.3.2 has been reproduced from Photochemical & Photobiological Sciences; Reference:
Labas et al., 2009, “Water disinfection with UVC radiation and H 2 O2 . A comparative study”.
Reproduced by permission of The Royal Society of Chemistry: www. res.org/pbs.
Graphics 15.7.1 has been reproduced from Photochemical & Photobiological Sciences; Refer-
ence: Labas et al., 2009, “Water disinfection with UVC radiation and H 2 O2 . A comparative study”.
Reproduced by permission of The Royal Society of Chemistry: www. res.org/pbs.
Graphics 15.7.2. has been reproduced from Photochemical & Photobiological Sciences; Ref-
erence: Labas et al., 2009, “Water disinfection with UVC radiation and H 2 O2 . A comparative
study”. Reproduced by permission of The Royal Society of Chemistry: www. res.org/pbs.

APPENDIX

Emission model for tubular germicidal lamps


In the integral of Equation (15.9) (Section 15.3), the region of integration for the fluence rate may
be divided into two parts: (i) the first, corresponding to the solid angle delimited by the lamp
boundaries (s) that is the solid angle that transports radiative energy coming from the body of
the lamp, and (ii) the solid angle corresponding to the remaining of the surrounding space, then:
⎡ ⎤
 
Eλ,0 = ⎣ Lλ, (x, t)d + Lλ, (x, t)d⎦ (15.A-1)
s 4π −s

Obviously, no radiation comes from the space defined by the solid angle 4π − s and the only
non-zero integral is the first. The final result can be written in terms of a spherical coordinate
system with origin at the point of incidence I (See Figs. 15.A-1a,b); then:
⎡ ⎤
 φ2  θ2  ,ϕ)
s̄=s(x,θ
⎢ ⎥
Eλ,o (x,t) = dφ dθ sin θ Lλ, (θ , φ,t) exp⎣− κs (t,s̄)d s̄⎦ (15.A-2)
φ1 θ1
s̄=sR (θ ,ϕ)

Two important observations must be made in connection with Equation (15.A-2):


1. After writing the equation in terms of the spherical coordinate system, the solid angle corre-
sponding to the space defined by the lamp boundaries (s ) originated a first integration in θ (φ)
and a second one in φ. These integrals have limits of integration that are complex functions of
the system geometry. The methodology for obtaining these limits will be illustrated following
an approach proposed by Irazoqui et al. (1973). For the moment we can say (Fig. 15A-1) that
these angles define, in a spherical coordinate system which is the natural system for radiation
propagation, the volume of the lamp that is able to send light to the chosen point inside the
reactor, the point x at s = s.
2. In Equations (15.A-1) and (15.A-2) the boundary condition has taken the form of L0λ (φ,θ,t).
This value must be provided by a lamp emission model. At this point it suffices to say that:
 
System geometry
L0λ (θ , f , t) = f (15.A-3)
Lamp characteristics

Equation (15.A-2) provides, in mathematical form, all the elements needed to compute the flu-
ence rate in homogeneous media. With an emission model for the lamp and the proper integration
limits for the lamp radiation contributions to an arbitrary point of incidence (In ), it is the possible
274 M. Flores et al.

Figure 15.A1. (a) Limits of integration for the source, (b) The extended, voluminal, isotropic emission
source model (adapted from Cassano et al., 1995).

to calculate the local value of the radiation energy absorbed per unit time and unit reaction volume
at any point of the reactor. For monochromatic radiation, this is the value required to formulate
the kinetics expressed by Equations (15.3) to (15.5).
As indicated before, the modeling of the lamp emission should provide the boundary condition
for the radiative transfer equation inside the reactor. There are lamps that produce an arc that emits
radiation by itself and, consequently, photons come out directly from such an arc. The whole lamp
volume makes emission. For example, this is the case of mercury arc low, medium and high-
pressure tubular lamps. We speak in these cases of “voluminal emission”. Voluminal emission
may be safely modeled as an isotropic emission; in this case the spectral radiance associated with
each bundle of radiation originated in some element of volume of the lamp is independent of
direction, and the associated emitted energy (per unit time and unit area) is also isotropic.
The following assumptions are made (Cassano et al. (1995)):

1. The emitters of the radiation source are uniformly distributed over the region of emission (a
volume).
2. In terms of Specific Intensities each elementary extension of emission has an isotropic emission
but the outgoing radiation energy is also isotropic when the emitting element is a volume.
3. Any emission element of the lamp emits per unit time, and for a given wavelength, an amount
of energy proportional to its extension and independent of its position inside the lamp volume.
4. When emission is voluminal, each of the differential volumes of emission is transparent to the
emission of its surroundings (an approximation which is not important if one works with the
information about the lamp output).
5. The lamp is a perfect cylinder bounded by mathematical surfaces with zero thickness. Hence,
any bundle of radiation coming from inside does not change its intensity or direction when
it crosses this boundary (an approximation which is not important if one works with the
information about the lamp output).
6. The lamp is long enough; consequently, neglecting end effects, the emission produced by the
lamp along its central axis is uniform. This assumption does not impose uniformity on the
Water disinfection with UVC and/or chemical inactivation 275

radiation field generated along the direction of the central axis. All what is said here is that
“end effects” are neglected.
7. Emission from the lamp is at steady state.
8. The lamp has a length LL and a radius rL.
9. A spherical coordinate systems located at each point of radiation reception (In) inside the
reactor can characterize the arriving Specific Intensity (also called radiance). It is necessary
to know the distance from such a point to the centerline of the lamp and two pairs of angular
coordinates [(θ 1, θ 2), (φ1,φ2)] that define the extension of the useful emitting volume of the
lamp.
Since a volume produces emission, the general radiative transfer equation (Alfano and Cassano,
2009) can be applied inside the lamp. There is no absorption (assumption 4) and no scattering.
The resulting equation is:
dLλ, (s, t)
= jλ,
e
(s, t) (15.A-4)
ds
Here, s is a directional coordinate in a three-dimensional space. From assumptions 2, 3 and 6,
emission is isotropic (independent of direction) and uniform (independent of position). Hence, at
steady state:
Lλ (0) = 0 (15.A-5)
In the lamp, along the direction , at s = 0, there is no entrance of radiation; this situation
provides the required boundary condition for Equation (15.8):
s= 0 Lλ (0) = 0 (15.A-6)

Integrating from s = 0 to s = ss (see Fig. 15.A-1) and using the following change of coordinates:
ds = −dρ (15.A-7)

where ρ corresponds to the spherical coordinate system defined at point In .


s=0 ρ = ρ2 (15.A-8)

s = ss ρ = ρ1 (15.A-9)
ρ1 and ρ2 are defined in Figure 15.A-1. Then, one gets:
Lλ (ss ) = jλe [ρs (x, θ , φ)] (15.A-10)

with:
[ρs (x, θ, φ)] = ρ2 (x, θ , φ) − ρ1 (x, θ , φ) (15.A-11)
It must be remarked that, as it should have been expected, ρs is a function of the position x in
the reactor and the direction of the incoming radiation given by the spherical coordinates (θ , φ).
Once more, from s = ss to s = sR there is no emission, no scattering and no absorption (the
medium has been assumed to be diactinic); from assumption 5 there is no refraction or reflection
at the lamp boundaries, then:
dLλ (s)
=0 (15.A-12)
ds
Consequently:
L0λ (θ, ϕ) = Lλ (sR ) = Lλ (ss )ϒR,λ (15.A-13)
Finally, the boundary condition is:
L0λ (θ, ϕ) = jλe [ρs (r, θ , ϕ)]ϒR,λ (15.A-14)

Reflection and absorption at the reactor wall were accounted for by means of the wall
transmission coefficient ϒR,λ .
276 M. Flores et al.

The value of jλe must be related to the monochromatic lamp output power Pλ,s . By definition,
jλe is the energy emitted per unit volume, unit solid angle of emission and unit time; then:
dPλ,s = jλe dVe d (15.A-15)
 
Pλ,s = jλe dVe d (15.A-16)
S VS

Since jλe corresponds to an isotropic and uniform emission:


Pλ,s Pλ,s Pλ,s
jλe = 1 1 = 2
= 2 2 (15.A-17)
VS dVe S d 4π (π rL LL ) 4π rL LL

From Equations (15.A-11) and (15.A-14) we must know the values of ρ2 (x, θ , φ) and ρ1 (x, θ , φ).
In order to know these values one must obtain explicit expressions for the independent variable
ρ, at the positions indicated by Equations (15.A-7) and (15.A-8). In the case of our reactor they
are seen in Figure 15.A-1. Let us consider a point located in an arbitrary position In , having
coordinates x(r, z, β). Let us look at an arbitrary direction (θ, φ).
Using the nomenclature indicated in Figure 15.A-1, from standard analytic geometry, the
equation of the boundary surface of the radiation source (a cylinder) in spherical coordinates is
written as follows:
ρ 2 sin2 θ − 2ρ(sin θ cos φ)r + (r 2 − rL2 ) = 0 (15.A-18)
The two solutions of this quadratic equation are precisely the values of ρ; i.e., they are the
intersections of the ρ coordinate with the front and rear parts of the lamp at any value of θ and φ:
r cos φ ± (r 2 cos2 φ − r 2 + rL2 )½
ρ1,2 = (15.A-19)
sin θ
Finally, the following value for ρS is obtained:
½
2[r 2 (cos2 ϕ − 1) + rL2 ]
ρs = (15.A-20)
sin θ
Equations (15.A-14), (15.A-17) and (15.A-20) provide the boundary condition for the emission
model, i.e.:
½
Pλ,S ϒR,λ [r 2 (cos2 ϕ − 1) + rL2 ]
L0λ (x, θ, ϕ) = 2
(15.A-21)
2π 2 rL LL sin θ
It must be noticed that either in Equation (15.A-21), all the characteristics of the lamp and
the relative position of the reactor with respect to the lamp are included. When this value is
used to calculate the photon absorption rate that is required to formulate the reaction rate inside
the reactor, this information is fully incorporated into the mass balances, i.e., we have the lamp
characteristics and its position inside the reactor incorporated as part of the design methodology.
Limits of integration for the variables θ and φ
The spectral radiance is a function of the directional coordinates. The limits of integration for the
independent variables θ and φ must be obtained.
Limits for the independent variable θ
Limiting rays coming from the lamp and reaching the generic point (r, β, z) inside the reactor
must satisfy two conditions:
(i) if the ray limits the value of θ , its equation (a straight line in space) must have a common
solution with the equation of the circumference that define the opaque zone (or lamp ends) of
the upper and lower parts of the lamp; however, for any plane at constant φ (a plane in r−z)
there are two values of θ that satisfy this condition for both the upper and lower boundaries;
Water disinfection with UVC and/or chemical inactivation 277

(ii) to eliminate this ambiguity a second restriction on θ must be imposed: one must choose the
angle corresponding to the intersection of the ray with that portion of the circumference that,
limited by the two generatrix lines corresponding to the limiting angles of φ, is closer to the
generic point (r, β, z), as indicated in the figure.
From Figure 15.A-1:
(LL − z) = ρ1 cos θ1 (15.A-22)
−z = ρ1 cos θ2 (15.A-23)
and with Equation (15.A-19):
 ½

−1 r cos φ − [r 2 (cos2 φ − 1) + rL2 ]
θ1 (φ) = tan (15.A-24)
(LL − z)
 ½

−1 r cos φ − [r 2 (cos2 φ − 1) + rL2 ]
θ2 (φ) = tan (15.A-25)
−z
Limits for the independent variable φ
The limiting rays in the φ-direction, for any value of the angle θ, must be tangent to the lamp
boundary at points located on the two generatrix lines of the cylinder (see Fig. 15.A-1). These
values can be obtained by imposing a restriction in the values of ρ 1 and ρ 2. At the limiting points,
since ρ = 0, both intersections of the ρ-coordinate with the lamp boundary must coincide,
i.e., in a projection of the lamp on the x-y plane, the limiting rays are tangent to the directrix
circumference of the lamp. This means that the condition in space is:
ρ1 = ρ2 (15.A-26)

From Equation (15.A-19) the following equation is obtained:


r 2 cos2 φ = r 2 − rL2 (15.A-27)

and since φ can only take on values in the first and fourth (negative value) quadrants, we have:
 
−1 (r − rL )
2 2 ½
−φ1 = φ2 = cos (15.A-28)
r

REFERENCES

Ahmed, Abd El-Shafey, I., Cavalli, G., Bushell, M.E., Wardell, J. N., Pedley, S., Charles, K. & Hay, N.J.: New
approach to produce water free of bacteria, viruses, and halogens in a recyclable system. Appl. Environ.
Microbiol. 77:3 (2011), pp. 847–853.
Alfano, O.M. & Cassano, A.E.: Scaling-up photoreactors. Applications to advanced oxidation processes. In:
H. De Lasa & B. Serrano Rosales (eds): Advances in chemical engineering. Elsevier. New York. 2009,
pp. 229–287.
Alkam, U., Teksoy, A. Atesli, A. & Baskaya, H.: Efficiency of the UV/H2 O2 process for the disinfection of
humic surface waters. J. Environ. Sci. Health A 42:4 (2007), pp. 497–506.
APHA: Compendium of Methods for the microbiological examination of Foods, 2nd ed, in M.L Speck (ed).
American Public Health Association, Washington, D.C., 1984.
Baldry, M.G.C.: The bactericidal, fungicidal, and sporicidal properties of hydrogen peroxide and peracetic
acid. J. Appl. Bacteriol. 54 (1983), pp. 417– 423.
Baldry, M.G.C. & Fraser, J.A.L.: Disinfection with peroxygens. In: K.R. Payne KR (ed): Industrial biocides.
Wiley, New York, 1988, pp. 91–116.
Baldry, M,G.C. & French, M.S.: Disinfection of sewage effluent with peracetic acid. Water Sci. Technol. 21
(1989b), pp. 203–206.
Bayliss, C. & Waites, W.: The combined effect of hydrogen peroxide and ultraviolet irradiation on bacterial
spores. J. Appl. Bacteriol. 47 (1979), pp. 263–269.
278 M. Flores et al.

Bianchini, L., Carlucci, L, Caretti, C., Lubello C., Pinzino, C. & Piscitelli, M.: An EPR study of wastewater
disinfection by peracetic acid, hydrogen perocide and UV. An. Chim. 92:9 (2002), pp. 783–793.
Block, S.S.: Disinfection, sterilization, and preservation. 4th ed. Lea & Febiger Pubs, Philadelphia, 1991,
pp. 167–181.
Bolton, J.R. & Cotton, C.A.: Ultraviolet disinfection handbook. American Water Works Association, Denver,
CO, 2008.
Caretti, C. & Lubello, C.: Wastewater disinfection with PAA and UV combined treatment: A pilot plant
study. Wat. Res. 37 (2003), pp. 2365–2371.
Cassano, A.E., Martin, C.A., Brandi, R.J. & Alfano, O.M.: Photoreactor analysis and design fundaments and
aplications. Ind.Eng. Chem. Res. 34 (1995), pp. 2155–2201.
Cerf, O.: Tailing of survival curves of bacterial spores. J. Appl. Bacteriol. 42 (1977), pp. 1–19.
Chick, H.: An investigation of the laws of disinfection. J. Hygiene 8 (1908), pp. 92–157.
Ciotti, C., Baciocchi, R., Cleriti, B. & Chiavola, A.: Peroxy-acids as an innovative oxidant for the remediation
of contaminanted sediment. 10th International UFZ/TNO Conference on Soil Water systems (Consoil),
2008.
Cords, B.R., Burnett, S.L., Hilgren, J., Finley, M., & Magnuson, J.: Sanitizers: Halogens, surface-active
agents, and peroxides. In: P.M. Davidson, J.N. Sofos & A.L. Branen (eds): Antimicrobials in food. 3rd
ed., CRC Press Taylor & Francis Group, FL, 2005.
Dalrymple, O.K., Stefanalos, E., Trotz, M.A. & Goswami, D.Y.: A review of the mechanisms and modeling
of photocatalytic disinfection. App. Catal. B 98 (2010), pp. 27–38.
El-Agamey, A. & McGarvey, D.J.: Evidence for a lack o reactivity of carotenoid addition radicals towards
oxygen: A laser flash photolysis study of the reaction of carotenoid with acylperoxyl radicals in polar and
non-polar-solvents. J. Am.Chem. Soc. 125 (2003), pp. 3330–3340.
Flores, M.J., Brandi, R.J., Cassano, A.E. & Labas, M.D.: Chemical disinfection with H2 O2 . The proposal or
a reaction kinetic model. Chem. Eng. J. 198–199 (2012), pp. 388–396.
Fraser, J.A.L., Godfree, A,F. & Jones, F.: Use of peracetic acid in operational sewage sludge disponsal to
pasture. Water Sci. Technol. 17 (1984), pp. 451–466.
Gard, S.: Chemical Inactivation of viruses. In: G.E.W. Wolstenholme & E.C.P. Millar (eds): Proceedings of
CIBA Foundation Symposium on Nature of Viruses. Churchill, London, UK, 1957, pp. 123–146.
Gardner, D.W.M. & Shama, G.: The kinetics of Bacillus subtilis spore inactivation on filter paper by UV
light and UV light in combination with hydrogen peroxide. J. Appl. Microbiol. 84 (1992), pp. 633–641.
Haas, Ch.: Disinfection, Chapter 14, Water Quality & Treatment. American Water Works Association
(AWWA). Fifth Edition, McGraw-Hill Inc., 1999.
Häder, D.P. & Sinha, R.P.: Solar ultraviolet radiation-induced DNA damage in aquatic organisms: Potential
environment impact. Mutat. Res. 571 (2005), pp. 221–233.
Harm, W.: Biological effects of ultraviolet radiation. Cambridge University Press, Cambridge, 1980.
Heywood, D. & Phillips, B., Stansbury, J.H.: Communications. Free radical hydroxylations with peracetic
acid. J. Org. Chem. 26 (1961), pp. 281–281.
Hom, L.W.: Kinetics of chlorine disinfection in an ecosystem. J. of the Environ. Division of the ASCE 98
(1972), pp. 183 -194.
Irazoqui, H.A., Cerda, J. & Cassano, A.E.: Radiation profiles in an empty anular photoreactor with a source
of finite spatial dimensions. AIChe J. 19 (1973), pp. 460–467.
Keller, B.K., Wojcik, M.D. & Fletcher, T.R.: A directly-disociative stepwise reaction mechanism for gas-phase
peroxyacetic acid. J. Photochem. Photobiolog. A: Chem. 195 (2008), pp. 10–22.
Kehrer, J. P.: The Haber-Weiss reaction and mechanisms of toxicity. Toxicology 692:149 (2000),
pp. 43–50.
Koubet, E.: The kinetics and mechanism of the decomposition of aliphatic peroxyacid in aqueous solutions.
PhD thesis, Brown University, New York, 1964.
Labas, M.D., Martin, C.A. & Cassano, A.E.: Kinetics of bacteria disinfection with UV radiation in an
absorbing and nutritious medium. Chem. Eng. J. 114 (2005), pp. 87–97.
Labas, M.D., Brandi, R.J., Martin, C.A. & Cassano, A.E.: Kinetics of bacteria inactivation employing UV
radiation under clear water conditions. Chem. Eng. J. 121 (2006), pp. 135–145.
Labas, M.D., Zalazar, C.S., Brandi, R.J. & Cassano, A.E. : Reaction kinetics of bacteria disinfection
employing hydrogen peroxide. Biochem. Eng. J. 38 (2008), pp. 78–87.
Labas, M.D., Brandi, R.J., Zalazar, C.S. & Cassano, A.E.: Water disinfection with UVC radiation and H2 O2 ,
A comparative study. Photochem. Photobiol. Sci. 8 (2009), pp. 670–676.
Leaper, S.: Influence of temperature on the synergistic sporicidal effect of peracetic acid plus hydrogen
peroxide on Bacillus subtilis SA22 (NCA 72–52). Food Microbiol. 1 (1984), pp. 199–203.
Water disinfection with UVC and/or chemical inactivation 279

Maillard, J.Y.: Bactericidal target sites for biocide action. J. Appl. Microbiol. Symposium Supplement. 92
(2002), pp. 16–27.
Marshall, R.: Standard methods for the examination of dairy products, 16th ed., Editorial APHA, American
Public Health Association, Washington D.C., 1992.
McDonnell, G. & Russell, A.D.: Antiseptics and disinfectants: Activity, action, and resistance. Clin.
Microbiol. Rev. 12:1 (1999), pp. 147–179.
Murov, S.L., Carmichel, I. & Hug, G.L.: Handbook of photochemistry: Potassium ferrioxalate actinometry.
2nd ed, Marcel Dekker Inc., New York, 1993.
Raffellini, S., Schenk, M., Guerrero, S. & Alzamora, S.M.: Kinetics of Escherichia coli inactivation employ-
ing hydrogen peroxide at varying temperatures, pH and concentrations. Food Control 22 (2011), pp.
920–932.
Rajala-Mustonen, R. & Heinomen-Tanski, H.: Effect of advancedoxidation processes on inactivation of
coliphages. Water Sci. Technol. 31(1995), pp. 131–134.
Rincon, A. & Pulgarin, C.: Effect of pH, inorganic ions, organic matter and H2 O2 on Escherichia coli K12
photocatalytic inactivation by TiO2 . Implications in solar water disinfection. Appl. Catal. B: Environ. 51
(2004), pp. 283–302.
Rokhina, E.V., Makarova, K., Golovina, E.A., Van As, H. & Virkutyte, J.: Free radical reaction pathway
thermochemistry of peracetic acid homolysis, and its application for phenol degradation: Spectroscopic
study and quantum chemistry calculations. Environ. Sci. Technol. 44 (2010), pp. 6815–6821.
Selleck, R.E., Saunier, B.M. & Collins, H.F.: Kinetic of bacteria deactivation with chlorine. J. Environm.
Eng. 104 (1978), pp. 1197–1212.
Severin, B. Kinetic modeling of microbial inactivation by ultraviolet. PhD Thesis, University of Illinois at
Urbana-Champaign, IL, 1982.
Severin, B., Suidan, M. & Engelbrecht, R.: Kinetic modeling of U.V. disinfection of water. Water Res. 17
(1983), pp. 1669–1678.
Shi, H.C. & Li,Y.: Formation of nitroxide radicals from secondary amines and peracids: A peroxyl radical oxi-
dation pathway derived from electron spin resonance detection and density functional theory calculation.
J. Mol. Catal. A. 271 (2007), pp. 32–41.
Sies, H.: Role of the reactive oxygen species in biological process, Klin. 690 Wochenschr. 69 (1991) 965–968.
Smith, G.: UV disinfection explained. UV disinfection & validation. WaterWork 2011, pp. 8–12.
Standard, J., Abbis, S. & Wood, J.M.: Combined treatment with hydrogen peroxide and ultraviolet irradiation
to reduce microbial contamination levels in pre-formed food packaging cartons. J. Food Protect. 46
(1983), pp. 1060–1064.
Sundstrom, D.W., Wier, B.A., Barber T.A. & Klein, H.E.: Destruction of pollutants and microorganisms in
water by UV light and hydrogen peroxide. Water Pollut. Res. J. Can. 27 (1992), pp. 57–68.
Trussell, R.R. & Chao, J.-L.: Rational design of chlorine contact facilities. J. Water Pollut. Control Federation
49:7 (1977), pp. 659–667.
Tutumi, M., Imamura, K., Hatano, S. & Watanabe, T.: Antimicrobial action of peracetic acid. J. Food Hyg.
Soc. 15 (1973), pp. 116–120.
US EPA: Drinking water, national primary drinking water regulation, final rules. Fed. Regisr. 54: 27486–
27541. U.S. Environmental Protection Agency, 1989.
US EPA: Small drinking water systems handbook: A guide to “packaged” filtration and disinfection
technologies with remote monitoring and control tools. U.S. Environmental Protection Agency, 2003.
US EPA: Ultraviolet disinfection guidance manual for the final long term 2 Enhanced Surface WaterTreatment
Rule. Office of Water (4601), EPA 815-R-06-007, 2006.
Watson, H.E.: A note on the variation of the rate of disinfection with change in the concentration of the
disinfectant. J. Hyg. 8 (1908), pp. 536–542.
Weber, S.: Light-driven enzymatic catalysis of DNA repair: A review of recent biophysical studies on
photolyase. BBA-Bioenergetics 1707:1 (2005), pp. 1–23.
Yuan, Z., Ni, Y. & Van Heiningen A.R.P.: Kinetics of peracetic acid decomposition. Part I. Spontaneous
decomposition at typical pulp bleanching condition. Can. J. Chem. Eng. 75 (1997), pp. 37–41.
Zalazar, C.S., Labas, M.D., Martin C.A., Brandi, R.J., Alfano, O.M. & Cassano, A.E.: The extended used of
actinometry in the interpretation of photochemical reaction engineering data. Chem. Eng. J. 109 (2005),
pp. 67–81.
This page intentionally left blank
CHAPTER 16

Ag/AgCl composite material: synthesis, characterization and


application in treating wastewater

Wei-Lin Dai, Quan-Jing Zhu, Jian-Feng Guo & Bo-Wen Ma

16.1 INTRODUCTION

Up to now, many kinds of semiconductor metal oxides have been widely studied. Among them,
TiO2 has been considered as a promising approach to solving environmental and energy problems
(Davis and Green, 1999; Du et al., 2011; Fujishima and Honda, 1972; Fox and Dulay, 1993;
Fujishima et al., 1994; Hurum et al., 2005; O’Regan and Grätzel, 1991; Tang et al., 2004; Wang
et al., 2012), owing to its ability to split water under ultraviolet (UV) light, and its potential
in other photoelectron-chemical solar-energy conversion with photochemical stability, non-toxic
nature and low cost (O’Regan and Graetzel, 1991). However, the overall efficiency is largely
inhibited under sunlight which consists of 43% visible and only 5% UV fraction, due to its wide
band gap (3.2 eV for anatase and 3.0 eV for rutile), which nonetheless permits only UV light
to be used (Hoffmann et al., 1995; O’Regan and Graetzel, 1991). Thus, it is of significance to
develop a photocatalyst that can harness visible light with high efficiency under normal sunlight
conditions.
During these years, much effort has been devoted to enhancing the photocatalytic efficiency and
visible-light utilization of TiO2 , which include sensitization (Abe et al., 2002; Sato et al., 2005),
impurity doping (Haick and Paz, 2003; Lu et al., 2003) and coupling with other semiconductors
(Bessekhouad et al., 2006; Wang et al., 2002). Composite semiconductors have been reported to be
potential as photocatalysts due to the fact that they can reduce the recombination of photogenerated
electrons and holes (Chen et al., 2005; Yu et al., 2007), and therefore can prolong the lifetime of
the charge carriers and enhance the quantum yield (Lee and Shih, 2007).
Noble metal nanoparticles (NPs) exhibit characteristic optical and physical properties that are
substantially different from those of the corresponding bulk material. In particular, silver NPs
show efficient plasmon resonance in the visible region. On the basis of this character, Awazu
et al. (2008) developed a method for preparing plasmonic photocatalyst, during which TiO2
was deposited on the surface of SiO2 @Ag NPs to prevent the oxidation of Ag by direct contact
with TiO2 . Recently, Wang et al. (2008; 2009) have synthesized Ag@AgCl and Ag@AgBr
plasmonic photocatalysts via an ion exchange and photoreduction method, which showed excellent
photocatalytic activity under visible light irradiation. It is reported that silver halide, generally
considered as photographic films, can coexist stably with Ag NPs during the whole photocatalytic
reaction process. It is also observed by Kakuta et al. (1999) that Ag0 species could contribute
to the smooth separation of electron-hole on the AgBr/SiO2 NPs and enable it to catalyze H2
production from alcohol radicals. After that, a series of such kind of plasmonic photocatalyst has
been prepared and investigated. For instance, Hu et al. (2010) synthesized a composite material
with Ag-AgI supported on mesoporous alumina (Ag-AgI/Al2 O3 ) and put forth the mechanism of
plasmon-induced photocatalytic reaction.
The cation dye of rhodamine B (RhB), well known for its good stability, is one kind of commonly
used xanthenes dyes for textile industry. The RhB can well dissolve in water or organic solvent and
has been found to be potentially toxic and carcinogenic, which make it be banned from application
in cosmetics and foods. Likewise, 4-chlorophenol (4-CP) can have serious consequences for
281
282 W.-L. Dai et al.

AgNO3

citric acid

ascorbic acid
FeC13
FeC13
+e–

Ag sphere Ag sphere AgC1 particles Ag/AgC1 core-shell

Scheme 16.1. Schematic description of the preparation of the core-shell Ag/AgCl sphere material (Ma
et al., 2013).

human health and the quality of the environment, particularly when released into natural water.
Thus, RhB and 4-CP are usually used as target pollutants in the photocatalyst research about the
purification of practical industrial waste water. In addition, series of industrial activities have
disturbed the geological equilibrium of metal ions, such as Hg(II), Pb(II), Cd(II), Ag(I), Ni(II),
and Cr(VI) ions, which act as the hazardous pollutants, through the release of large quantities of
toxic metal ions into the environment.
Herein, we prepared several kinds of Ag/AgCl composite materials, including Ag/AgCl core-
shell sphere (Ma et al., 2013), Ag/AgCl@Cotton-fabric (Ma et al., 2011), Ag-AgCl/WO3 hollow
sphere (Ma et al., 2012) and Ag-AgCl@TiO2 (Guo et al., 2012), and Ag-AgI/Fe3 O4 @SiO2
(Guo et al., 2011) through simple and generic approaches. RhB, 4-CP, MO and/or Cr(VI) ion
were chosen as the target pollutants to investigate the photocatalytic activity of the catalysts
under visible light irradiation. N2 adsorption and desorption isotherm, X-ray diffraction, X-ray
photoelectron spectroscopy, transmission electron microscopy, or scanning electron microscopy
were used to determine the correlation between the micro-structure and the catalytic properties
of the as-prepared photocatalysts. A plasmon induced photocatalytic mechanism has also been
verified.

16.2 SYNTHESIS OF THE PHOTOCATALYSTS

16.2.1 Ag/AgCl core-shell sphere


All chemicals were of analytical reagent grade and purchased from Sinopharm Chemical Reagent
Co. Ltd. (SCRC), and used without further purification. The synthesis process is shown as
Scheme 16.1.

16.2.1.1 Preparation of Ag spheres using ascorbic acid as the reducing agent


AgNO3 aqueous solution and citric acid aqueous solution were added to 100 mL of deionized
water in a 250 mL beaker with a magnetic stirrer in an ice–water bath. An ascorbic acid aqueous
solution was then quickly injected into the vigorously stirred mixture. The added ascorbic acid was
kept at equal molar ratio to that of silver ions. The solution became khaki-colored immediately,
and a large quantity of Ag particles was produced in a few minutes. After 15 min, the particles
were collected by centrifugation and repeatedly rinsed with deionized water. Ag spheres were
obtained after the samples were dried at 60◦ C for 5 h.
Ag/AgCl composite material 283

16.2.1.2 Preparation of Ag/AgCl core-shell sphere using ferric chloride


In a typical Ag/AgCl core-shell sphere synthesis, the reaction mixture containing the as-
synthesized Ag spheres solution was added to an aqueous solution containing 50 mM PVP. Then,
a certain amount of FeCl3 (the molar ratio of Fe:Ag = 1:10, 1:5, 1:2, 1:3, 1:4, 1:1 and 2:1) was
slowly added dropwise to this Ag spheres solution. The resulting mixture was maintained at room
temperature until its color turned brown. Vigorous stirring was employed throughout the synthe-
sis. The as-obtained samples for morphology and structure analysis were washed with water and
ethanol to remove the FeCl3 and PVP via centrifugation. Finally, pure Ag/AgCl core-shell spheres
were obtained.

16.2.2 Ag/AgCl@Cotton-fabric
Before use, the cotton fabric was cleaned with deionized water at 100◦ C and dried in an oven at
80◦ C. Firstly, 0.09 g of AgNO3 and 10 mL of deionized water were added to a 100 mL of beaker
and were stirred for 10 min. Then the clean cotton fabric was put into this solution. After that, it
was heated at 80◦ C under stirring until all of the solvent was almost evaporated. Secondly, after
dried at 80◦ C, the cotton fabric was immersed into 0.5 M KCl solution for 5 h, and thus the AgCl
NPs were deposited on the surface of the cotton fabric. Finally, the as-prepared samples were
irradiated with 4 × 8 W ultraviolet lamps for a given time to reduce the partial Ag+ ions in the
AgCl particles to Ag0 species. Before use, the samples were dried at 80◦ C overnight.

16.2.3 Ag-AgCl/WO3 hollow sphere


16.2.3.1 Preparation of the hollow sphere PbWO4
The PbWO4 precursors were synthesized using a solvothermal process (Chen and Ye, 2008). In a
typical procedure, 1.51 g of Pb(AC)2 ·H2 O and 1.31 g of Na2WO4 were respectively dissolved in
two sets of 50 mL of aqueous ethylene glycol solution. Then the as-obtained solutions were mixed
under vigorous magnetic stirring at room temperature and white precipitates were generated. The
resulting suspension was further stirred for 10 min and then was transferred into a Teflon-lined
stainless steel autoclave, followed by heat treatment at 160◦ C for certain time. After cooling
to room temperature, the products was filtered, washed three times with alcohol/water mixture
(V/V = 1:1) and then dried in air at 80◦ C overnight.

16.2.3.2 Preparation of the hollow sphere WO3


To get hierarchical WO3 hollow structures, the PbWO4 precursors with different morphologies
were firstly immersed in 100 mL of 4 M HNO3 solution for 48 h. Then the products were filtered,
washed with distilled water, and dried in air. After then, the acid-treated products were calcined
in a furnace at 500◦ C for 2 h in air.

16.2.3.3 Preparation of Ag-AgCl/WO3


The photocatalyst were prepared by the deposition-precipitation-photoreduction method. 0.5 g of
as-prepared hollow sphere WO3 and 0.12 g of hexadecyltrimethylammonium chloride (CTAC)
were added into 100 mL of distilled water and the mixture was stirred at room temperature
for 30 min. The aqueous solution of AgNO3 was added into the above mixture. After being
stirred magnetically for 20 min, the above mixture was irradiated with a 4 × 8 W ultraviolet lamp
(λ = 254 nm) for certain minutes to reduce partial Ag+ ions in the AgCl particles to Ag0 species by
photochemical decomposition of AgCl. The as-prepared samples with different photoreduction
time were named as X Ag-AgCl/WO3 , where X stands for the irradiation time in minutes. After
that, the catalysts were collected, washed with distilled water and dried at 80◦ C for 24 h in dark.
Finally, the as-prepared products were calcined in air at 300◦ C for 3 h to obtain the Ag-AgCl/WO3
hollow sphere.
284 W.-L. Dai et al.

16.2.4 Ag-AgCl@TiO2
In a typical synthesis of Ag/AgCl@TiO2 , 0.2 g of commercial Degussa P25 TiO2 and 0.3 g of
CTAC were added to 100 mL of deionized water and the suspension was stirred for 60 min.
Then 2.0 mL of 0.1 M AgNO3 was quickly added to the above mixture. During this process,
the excessive surfactant CTAC not only adsorbed onto the surface of P25 to limit the number of
nucleation sites for AgCl to grow, resulting in homogenously dispersed AgCl, but also induced
Cl− to precipitate Ag+ in the suspension. The resulting suspension was stirred for 1.0 h and
then placed under irradiation of 4 × 8 W ultraviolet light for the indicated lengths of time. The
suspension was filtered, washed with deionized water, and dried at 80◦ C for 12 h. Then the
gray powder was calcined at 300◦ C for 3 h. Depending on the duration of irradiation, the as-
prepared catalysts were denoted as Ag-AgCl@TiO2 −m, where “m” represented 5, 10, 20, and
30 min of photo-reduction, respectively. For comparison, Ag@TiO2 was synthesized by a similar
process without CTAC. Nitrogen-doped N-TiO2 was also prepared as described previously for
comparison (Fang et al., 2007). First, 7 wt% of ammonia aqueous solution was added dropwise
to Ti(SO4 )2 aqueous solution to prepare Ti(OH)4 precipitate. After being filtrated and washed
with distilled water five times, the precipitate was dispersed into diluted HNO3 aqueous solution
(HNO3 /TiO2 (mole) = 0.6) at 60◦ C with stirring, and the precipitate was peptized and transformed
into a transparent sol with a pale blue tint. After being aged at 50◦ C, the sol became a gel. Finally,
the xerogel powder was obtained by drying the TiO2 gel at 50◦ C followed by grinding.

16.2.5 Ag-AgI/Fe3 O4 @SiO2


16.2.5.1 Synthesis of Fe3 O4 particles
The magnetic particles were prepared through a solvothermal reaction. Briefly, 6.0 g of
FeNO3 ·6H2 O, 3.0 g of PEG-20000 and 10.5 g of sodium acetate were dissolved in 100 mL of
ethylene glycol under magnetic stirring. The obtained homogeneous yellow solution was trans-
ferred to a Teflon-lined stainless-steel autoclave and sealed. After reacted at 200◦ C for 24 h, the
autoclave was cooled to room temperature. The obtained black magnetite particles were washed
with ethanol for 6 times, and then dried in vacuum at 60◦ C for 12 h.

16.2.5.2 Synthesis of Fe3 O4 @SiO2 microspheres


The core-shell Fe3 O4 @SiO2 microspheres were prepared as followings. Briefly, 0.10 g of Fe3 O4
particles (∼100–200 nm in diameter) were treated with 0.1 M HCl aqueous solution (50 mL)
by ultrasonication for 10 min. After that, the magnetite particles were separated and washed
with deionized water, and then homogeneously dispersed in the mixture of ethanol (80 mL),
deionized water (20 mL) and concentrated ammonia aqueous solution (1.0 mL, 25 wt%), followed
by the addition of TEOS (0.03 g, 0.144 mmol). After stirring at room temperature for 6 h, the
Fe3 O4 @SiO2 microspheres were separated and washed with ethanol and water for 6 times with
the assistance of a magnet. Finally, the purified microspheres of Fe3 O4 @ SiO2 were dried in
vacuum at 60◦ C for 12 h.

16.2.5.3 Synthesis of AgI/Fe3 O4 @SiO2


A suspension of Fe3 O4 @SiO2 (0.5 g) in 100 mL of deionized water was sonicated for 10 min.
KI (0.17 g) was then dispersed into the suspension, followed by another 30 min ultrasonication.
AgNO3 (0.17 g) in NH4 OH (2.0 mL, 25 wt%) was added to the mixture and stirred for another
12 h. The product (AgI/Fe3 O4 @SiO2 ) was filtered, washed with deionized water, and dried in air
at 70◦ C, then calcined in an inert atmosphere (Ar) at 500◦ C for 3 h.

16.2.5.4 Synthesis of Ag-AgI/Fe3 O4 @SiO2


Ag-AgI/Fe3 O4 @SiO2 was prepared via a photocatalytic reduction method. A suspension of
AgI/Fe3 O4 @SiO2 (0.5 g) in 50 mL of deionized water was sonicated for 10 min. AgNO3 (15 mg)
in NH4 OH (0.3 mL, 25 wt%) was then added to the mixture and stirred in the dark for 30 min.
Ag/AgCl composite material 285

Then the catalyst suspension was irradiated with ultraviolet light (4 × 8 W) (λ < 360 nm) for
different duration to reduce adsorbed Ag+ to Ag0 via AgNO3 photolysis. Finally, the product
was filtered, rinsed with deionized water, and dried at ambient temperature. On the basis of the
different irradiation duration, the as-prepared catalysts were denoted as 30Ag-AgI/Fe3 O4 @SiO2 ,
90Ag-AgI/Fe3 O4 @SiO2 and 150Ag-AgI/Fe3 O4 @SiO2 , corresponded to 30, 90 and 150 min pho-
toreduction time, respectively. Following this procedure, Ag/Fe3 O4 @SiO2 was prepared, except
that Fe3 O4 @SiO2 was replaced by AgI/Fe3 O4 @SiO2 and the photoreduction duration was 30 min.
As a reference, N-doping TiO2 was prepared via thermal annealing of TiO2 (P25, Degussa)
powder in pure NH3 flow at 500◦ C for 2 h, which was generally recognized as an efficient
visible-light-driven photocatalyst and denoted as N-TiO2 in this study (Chen et al., 2006).

16.3 CHARACTERIZATION OF THE PHOTOCATALYSTS

Diffuse reflectance spectroscopy was performed using a SHIMADZU UV-2450 instrument with a
collection speed of 40 nm min−1 using BaSO4 as the reference. Transmission electron micrographs
(TEM) were obtained using a JEOL 2011 microscope operating at accelerating voltage of 200 kV.
X-ray photoelectron spectroscopy (XPS) measurements were performed on a PHI 5000C ESCA
System with Mg Kα source at 14.0 kV and 25 mA. All the binding energies were referenced to
the contaminant C 1s peak at 284.6 eV of the surface adventitious carbon. Scanning electron
micrographs (SEM) were obtained using a PHILIPS XL 30 microscope operating at accelerating
voltage of 20 kV.

16.4 EVALUATION OF PHOTOCATALYTIC ACTIVITY

In a typical process, photocatalytic experiments were performed in a beaker placed under the lamp
bracket, containing aqueous suspensions of substrate 4-CP (100 mL, 10 mg L−1 ), RhB (100 mL,
50 mg L−1 ), MO (100 mL, 20 mg L−1 ) and 50 mg of catalyst powder. The light source was a 125 W
metal halide lamp (Philips) equipped with wavelength cutoff filters for λ ≤ 420 nm and focused
on the beaker. Prior to the irradiation, the suspension was ultrasonicated for 10 min and then
stirred in dark for 30 min to achieve the adsorption/desorption equilibrium. After turning on the
lamp, 2 mL of suspension was sampled at certain time intervals and centrifuged by centrifuge
(Shanghai Anting Scientific Instrument Factory, China) at 13,000 rpm for 10 min to remove the
particles. The upper clear liquid was analyzed by recording the characteristic absorption peak of
4-CP at 280 nm, RhB at 550 nm or MO at 463 nm to calculate the concentration of the compounds
according to the work curve previously drawn. The determination of Cr(VI) concentration was
performed by the 1,5-diphenylcarbazide (DPC) method (Wittbrodt and Palmer, 1995). In the
evaluation of activity for photo-reduction of Cr(VI), 0.2 mM of ethylenediaminetetraacetic acid,
disodium (EDTA) was used as sacrificial electron donator.

16.5 RESULTS AND DISCUSSION

16.5.1 Ag/AgCl core-shell sphere


SEM images of the Ag spheres and the Ag/AgCl core-shell spheres are showed in Figure 16.1
(A and B). It is obvious that the as-prepared Ag spheres are of uniform shape with an average
diameter of 1.0 µm. From Figure 16.1B, it can be seen that the perfect microspheres (1.0 µm in
diameter) are assembled by numerous closely packed Ag nanowires, which likely results from
the addition of citric acid. Thus, the surface of the Ag spheres prepared is rough and beneficial
to the oxidation reaction. SEM images of the Ag/AgCl core-shell spheres prepared by the Fe:Ag
ratio of 1:1 is presented in Figures 16.1C and 16.1D. It is interesting to find that the morphology
286 W.-L. Dai et al.

(A) (B)

10 µm 500 nm

(C) (D)

5 µm 1 µm

Ag
(E)
50
Intensity (a.u.)

40
30
CI
20
10
0
0 2 4 6 8 10
Energy (keV)

Figure 16.1. Typical SEM images of the synthesized Ag sphere (A,B), Ag/AgCl core-shell sphere (C,D)
and EDS of the Ag/AgCl core-shell sphere material (E) (Ma et al., 2013).

of the spheres was retained after the oxidation, although the diameter of the spheres increased to
2 µm. It can be seen that silver and chlorine elements were present in the samples from the EDS
spectrum of the Ag/AgCl core-shell spheres (Fig. 16.1E).
Figure 16.2 presents the SEM images of the Ag/AgCl core-shell spheres synthesized at various
molar ratios of Fe:Ag (1:10, 1:5, 1:1 and 2:1). From Figure 16.2J, it is confirmed that the core-
shell spheres were actually formed by assembled nanosheets arranged along different directions,
and it is found that the surface of the catalysts had changed from silver to silver chloride according
to the EDS spectrum results in Figure 16.1E. In Figure 16.2, it is observed that the morphology
of the materials varies with the change in the molar ratio. The surface of the spheres was formed
by assembled nanosheets of approximately uniform size and shape with distinct edges when the
ratio was as low as 1:10. With the increase of the amount of FeCl3 , the nanosheets grew in size,
and the surface of the Ag/AgCl core-shell spheres became increasingly smooth as the ratio was
increased to 1:5. When the ratio was 1:1, the outer spheres appeared to be tightly wrapped by a
layer of nanosheets which was obviously larger than that when the ratio was 1:5. This observation
implies that most silver on the surface may be completely etched and AgCl nanocrystals were
formed on the outer spheres. Whereas, the morphology of the sphere materials was destroyed
and the uniform size of the spheres disappeared when the ratio was further increased to 2:1.
Ag/AgCl composite material 287

Figure 16.2. Typical SEM images of Ag/AgCl-1:10 (A, B), Ag/AgCl-1:5 (C, D), and Ag/AgCl-1:1(E, F, I,
J), Ag/AgCl-2:1 (G, H) (Ma et al., 2013).

It is demonstrated that the growth of AgCl nanocrystals on the surface of catalysts was sufficient
and that there was no need to increase the ratio of Fe:Ag to 2:1. Therefore, during the synthesis
process, the optimal ratio of Fe:Ag is 1:1. Figures 16.2I and 16.2J show the core-shell structure
of the as-prepared Ag/AgCl spheres after being grounded in an agate mortar in the presence of
ethanol. It is also interesting to find that the Ag core was surrounded by AgCl shell, as shown in
the highlighted area.
288 W.-L. Dai et al.

Figure 16.3. (A) Typical UV-Vis DRS spectra of Ag sphere (a), Ag/AgCl-1:10 (b), Ag/AgCl-1:5 (c),
Ag/AgCl-1:1(d) and Ag/AgCl-2:1 (e); (B) Photocatalytic activities for the degradation of
RhB dye under visible-light irradiation (λ > 420 nm) without catalyst (a) and over various
Ag/AgCl core-shell spheres with different molar ratios of Fe:Ag (1:10 (b), 1:5 (c), 1:2 (d), 1:3
(e), 1:4 (f), 1:1 (g) and 2:1 (h)) (Ma et al., 2013).

The UV-vis DRS spectra of theAg sphere andAg/AgCl core-shell spheres synthesized at various
molar ratio of Fe:Ag (1:10, 1:5, 1:1 and 2:1) are shown in Figure 16.3A.All of the materials showed
strong absorption in the UV region, a typical characteristic of silver. Almost no absorption peak
in the 400–750 nm region was observed in the silver sphere sample, owing presumably to the
absence of surface plasmon resonance of Ag NPs. It is assumed that the silver particles made of
the silver sphere are not in nanoscale, and are too large to produce the surface plasmon absorption.
Nevertheless, the series of Ag/AgCl core-shell sphere photocatalysts exhibited broad absorption
in the visible light region of 400–750 nm, due to the surface plasmon resonance of Ag NPs. The
silver particles on the surface of spheres were etched by FeCl3 and silver nanoparticles were
formed during the oxidation reaction. According to the previous assumption, it was found that
the absorption band at about 500 nm, due to the surface plasmon resonance of nano-sized silver,
is proportional to the degree of oxidation. With increasing the amount of the oxidant, the amount
of Ag nanoparticles increased while their size decreased. When the wavelength of the irradiating
light was much larger than the diameter of silver NPs, the electromagnetic field across each entire
silver NP was essentially uniform. With oscillations in the electromagnetic field, the weakly bound
electrons of the silver nanoparticles respond collectively, leading to the plasmonic state. When
the incident light frequency matches the plasmonic oscillation frequency, the incident light is
absorbed, giving rise to surface plasmon absorption.
The photocatalytic activity of series of Ag/AgCl core-shell spheres was evaluated in the
photodegradation of RhB and MO in aqueous dispersions under visible light irradiation. The
photocatalytic activity of the Ag/AgCl core-shell spheres prepared by using different molar ratio
of Fe:Ag (1:10, 1:5, 1:2, 1:3, 1:4, 1:1 and 2:1) is shown in Figure 16.3B. As can be seen, at
irradiations of wavelengths of λ > 420 nm under visible light, the reaction showed limited photo-
catalytic activity for the degradation of RhB in the absence of any catalysts; but the presence of
Ag/AgCl core-shell sphere catalysts resulted in very high photodegradation activity. The Fe:Ag
(1:1) sample showed the most pronounced photocatalytic activity among the various Ag/AgCl
core-shell spheres catalysts, and it could completely decompose RhB dye solution after irradiation
for 20 min, which is similar to the reaction using Fe:Ag (2:1). Assuming that photodegradation of
RhB dye follows a pseudo first order reaction mechanism, the degradation rate of RhB dye in the
presence of catalysts prepared with the molar ratio of Fe:Ag as 1:10, 1:5, 1:2, 1:3, 1:4, 1:1 and
2:1 were estimated to be 0.079, 0.097, 0.123, 0.124, 0.165, 0.246, and 0.246 mg min−1 , respec-
tively. This result indicated that the photocatalytic activity of the samples was greatly enhanced by
increasing of the molar ratio of ferric chloride to Ag, and subsequent oxidation from Ag spheres
to AgCl crystal on the surface of the Ag spheres. In accord with previous reports (Bi and Ye, 2009;
Ag/AgCl composite material 289

Figure 16.4. Photodegradation of 4-CP in aqueous suspension under visible light irradiation over different
photocatalysts (a) blank; (b) Ag/AgCl core-shell spheres (1:1) (Ma et al., 2013).

Table 16.1. Surface composition of Ag/AgCl core-shell sphere prepared by different molar ratios of Fe:Ag
(Ma et al., 2013).

Sample FeCl3 : AgNO3 (M/M) Ag0 /Ag+1 (M/M) Ag/Cl 1) (M/M)

Ag/AgCl-1:10 1:10 2.31 2.10


Ag/AgCl-1:5 1:5 0.93 1.75
Ag/AgCl-1:2 1:2 0.79 1.12
Ag/AgCl-1:1 1:1 0.035 0.96
Ag/AgCl-2:1 2:1 ≈0 0.63

1 Molar ratio from XPS

Li et al., 2010), the photocatalytic activity highly depended on ratio of Ag:AgCl on the surface of
the catalysts. There should be an optimum ratio for the best photoactivity. The Ag-AgCl sample
prepared by a Fe:Ag (1:1) ratio may be considered as the best for optimum photoactivity in a fair
comparison of all the Ag/AgCl core-shell sphere samples. When the molar ratio was increased
from 1:1 to 2:1, the photodegradation activity was retained. Thus, we believe that the oxidant was
enough for the oxidation process when the molar ratio of Fe:Ag was 1:1. The photodegradation of
MO (not shown here) followed the similar trends to the RhB degradation in aqueous dispersions
under visible light irradiation, suggesting that the as-prepared core-shell spheres could serve as
a general photocatalyst. Additionally, in order to eliminate the possibility of self-degradation of
dyes (Mrowetz et al., 2004; Yan et al., 2006), the visible-light driven photocatalytic activity of the
as-prepared catalysts on the degradation of 4-CP was also evaluated. As shown in Figure 16.4,
it is interesting to find that after 30 minutes’ reaction, 4-CP was totally degraded, indicating the
superior photocatalytic performance of the Ag/AgCl core-shell spheres.
To further investigate the correlation between the ratio of Ag:AgCl on the surface of the
catalysts and the photocatalytic activity, the surface composition of Ag/AgCl core-shell spheres
was obtained from XPS measurements. As can be seen from Table 16.1, with increasing amount
of ferric chloride, the ratio of Ag0 :Ag+ decreased abruptly from 2.31 to approximately 0.0, while
the Ag:Cl ratio decreased from 2.1 to 0.63. It is believed that more and more Ag atoms were
oxidized to AgCl with increasing amount of ferric chloride, which is in accord with the results
from the UV-vis DRS of the catalysts. The ratio of Ag:AgCl, obtained in the sample prepared
by using Fe:Ag in (1:1) ratio, could be a good approach to achieve optimum photoactivity and
the optimum ratio of Ag0 :Ag+ was determined to be 0.035. This ratio appears to facilitate the
290 W.-L. Dai et al.

Figure 16.5. Schematic diagram illustrating the photogeneration of electrons/holes as well as their migration
and reaction in the surface of Ag/AgCl core-shell sphere (Ma et al., 2013).

interaction between Ag nanoparticles and AgCl crystals that results in the high photocatalytic
activity.
To understand the photocatalytic activity and the interaction betweenAg nanoparticles andAgCl
crystals, a possible mechanism was proposed in Figure 16.5, which is similar to that of Wang et al
(Wang et al., 2008). According to this mechanism, during the degradation process, the oxidation
reaction of H2 O takes place concurrently with the reduction of Ag+ , thus maintains stability
in the whole system. Additionally, the reduction of Cl and re-oxidation of Cl− happens during
the photocatalytic reactions, therefore, the whole process serves as a photocatalytic reaction.
Wang et al (Wang et al., 2008) found that the surface plasmon resonance of Ag NPs that results
in an absorption in the 400–750 nm visible light region may also prevent the recombination of
photoexcited holes and electrons, and meanwhile lead to combination of Cl− with the holes to form
active Cl0 . The Ag NPs on the surface would play a key role on the reaction. These Ag NPs would
include some Ag NPs that come from the decomposition of AgCl shell, and some Ag NPs that were
not oxidized. During the reaction, the surface of AgCl NPs becomes negatively charged because
Ag+ ions are reduced partially to Ag0 species by visible light irradiation. The whole process of the
degradation seems like a reaction cycle. In the present Ag/AgCl core-shell sphere system, when
Ag/AgCl core-shell sphere in the solution was irradiated by visible light, it is possible that electron-
hole pairs are photogenerated in Ag NPs due to SPR effect. Then the photoexcited electrons can be
transferred to the ubiquitous molecular oxygen to form the super-oxide radical O− 2 , subsequently
yielding the HOO• radicals through protonation; finally HO• radicals are also formed. All these
active radicals can contribute to the decomposition of RhB. Simultaneously, the surface of AgCl
NPs becomes negatively charged because Ag+ ions are reduced partially to Ag0 species by visible
light irradiation, and the electric field of AgCl is likely terminated by Cl− anions. Therefore,
the photoexcited holes are transferred to the surface of AgCl NPs and cause the oxidation of
Cl− anions to Cl0 atoms, which are subsequently able to oxidize RhB and to be reduced to Cl−
again. As a result, the Ag NPs can act as a sink for photoreduced charge carriers and be quickly
renewed; the Ag/AgCl core-shell sphere photocatalytic system can stay stable due to the chloride
reaction cycle. Further, if Ag NPs on the surface of the Ag/AgCl core-shell sphere are insufficient,
the enhancement of photocatalytic activity by the surface plasmon resonance would be limited.
If the Ag NPs were excessive, the energy levels may change to promote the recombination of
the electrons and holes, which would reduce the photocatalytic activity of Ag/AgCl core-shell
spheres. Additionally, if AgCl NPs were insufficient, the amount of Cl0 radicals may be greatly
reduced, which could also act as a factor in reducing the photocatalytic activity of samples. In
conclusion, at the optimum ratio of 0.035, the interaction between Ag nanoparticles and AgCl
leads to the best photocatalytic activity. This finding corroborates the mechanism proposed by
Hu et al. (Hu et al., 2010).
Ag/AgCl composite material 291

Figure 16.6. XRD patterns (A) of 0.0048 g/cm2 Ag/AgCl@cotton fabric with different photoreduction
duration (a) pure cotton fabric, (b) 0 min, (c) 3 min, (d) 5 min, (e) 10 min, and (f) 15 min; SEM
(B) and TEM (C) image of the as-prepared photocatalyst Ag/AgCl@cotton-fabric (Ma et al.,
2011).

Scheme 16.2. Schematic description of the process of photodegradation (Ma et al., 2011).

16.5.2 Ag/AgCl@Cotton-fabric
The XRD patterns of the support and 0.0048 g/cm2 Ag/AgCl@cotton-fabric with different pho-
toreduction duration is shown in Figure 16.6A. All samples can be seen the peaks of support
and KCl, which is not surprising because the samples were immersed in KCl solution for 5 h.
Diffraction peaks attributed to AgCl are observed for all the catalysts. As Ag NPs deposited on the
surface of the cotton fabric was not easy to be detected, no diffraction peaks assigned to Ag0 are
displayed in the sample with 3 min photoreduction. When the reduction time was increased to 5
min, the peaks ascribed to metallic Ag can be observed. It is indicated that different photoreduction
duration results in different ratio of Ag to AgCl.
The SEM image of the Ag/AgCl@cotton fabric product is presented in Fig. 16.6B. It is revealed
that AgCl particles with diameters in the range of 300 to 600 nm were deposited on the surface
of the cotton fabric. The TEM image of the as-prepared photocatalyst Ag/AgCl@cotton-fabric
indicated that the diameters of the Ag NPs are about 10 nm.
The photocatalytic activity of the samples was also evaluated by photodegradation of RhB
aqueous solution under visible light irradiation (Ma et al., 2011). In this experiment, the cotton
fabric was used as the support, some special devices was needed to carry out the photodegra-
dation. As can be seen in Scheme 16.2, the textile supporting Ag/AgCl was fixed on a Teflon
frame during the experiment, and then this frame was put into a beaker. The RhB aqueous solu-
tion was injected into the beaker before the photodegradation. After reaction, the solution was
extremely clear, which could be analyzed directly without any further treatments. Centrifugation
or filtration was no longer needed to obtain used catalyst, thus saving much energy. As shown
292 W.-L. Dai et al.

Figure 16.7. Photodegradation of RhB over 0.0048 g/cm2 Ag/AgCl@cotton-fabric with different photore-
duction duration: (a) 0 min, (b) 3 min, (c) 5 min, (d) 10 min, (e) 15 min, and (f) no catalyst.
Visible light λ > 420 nm (Ma et al., 2011).

Figure 16.8. Irradiation time dependence of the relative concentration of RhB over Ag/Ag@cotton-fabric
during repeated photodecomposition experiments under visible light irradiation (Ma et al.,
2011).

in Figure 16.7, the pure support showed no activity for the RhB photodegradation under visible
light (λ > 420 nm), which only caused a slight fading of RhB solution due to the absorption of a
small amount of RhB by the textile. Therefore, the dye self-photosensitization and the absorption
of the support with the dye could be ignored. Figure 16.7 presented that the sample with 3 min
photoreduction duration exhibits the best activity among these catalysts. It is known that differ-
ent photoreduction duration results in different ratio of Ag to AgCl. However, the sample with
0 min photoreduction duration exhibits a little lower activity than the sample with 3 min. It is
because some Ag+ ions on the sample with 0 min photoreduction duration can be reduced to Ag0
species under visible light irradiation during the photodegradation. The result also demonstrates
that there were certainly relations between the ratio of Ag to AgCl and the photocatalytic activity,
as it was illustrated in Ag/AgCl core-shell spheres before.
The durability of the Ag/AgCl @cotton fabric catalyst for the degradation of RhB under visible
light was shown in Figure 16.8. Due to its special structures, Ag/AgCl@cotton-fabric was easily
recycled without any treatment in these experiments. The photocatalytic activity in the degraded
RhB did not decrease significantly after six successive cycles under visible light irradiation,
demonstrating that Ag/AgCl@cotton-fabric is an effective and stable catalyst under visible light
Ag/AgCl composite material 293

Figure 16.9. SEM images of H2WO4 (A) and of 30Ag–AgCl/WO3 (B), (C) and (D) in different magnifica-
tion (Ma et al., 2012).

and UV irradiation. The stability of the Ag/AgCl@cotton-fabric photocatalyst under irradiation


may arise from the fact that a photon is absorbed by the silver NPs, and an electron separated from
an absorbed photo remains in the NPs rather than being transferred to the Ag+ ions of the AgCl
lattice (Wang et al., 2008). Generally, photogenerated electrons are expected to be trapped by O2
in the solution to form superoxide ions (O−2 ) and other reactive oxygen species (Soni et al., 2008).
Additionally, comparing with the powdery Ag/AgCl, the Ag/AgCl@cotton- fabric photocatalyst
has its distinctive advantage because the loss of the powdery Ag/AgCl would happen inevitably
after the treatment of recycling while the Ag/AgCl@cotton-fabric photocatalyst almost has no
loss during the recycling because Ag/AgCl was firmly fixed on the cotton fabric. Besides, we also
found that other organic pollutants such as phenol were also quickly degraded by this photocatalyst
under visible light irradiation.

16.5.3 Ag-AgCl/WO3 hollow sphere


Figure 16.9 showed the SEM images of these H2WO4 and 30Ag-AgCl/WO3 spheres calcined at
500◦ C. It is obvious that the samples kept the unique shape and the hollow sphere was not destroyed
before and after calcinations. Because of the structure of hollow sphere, the incident lights could
be reflected repeatedly and consequently enhanced the utilization efficiency of light, which might
improve the photocatalytic activity of the as-prepared catalysts. Figure 16.10 presented the TEM
images and EDS results of 30Ag-AgCl/WO3 photocatalyst. It can be seen that the Ag nanoparticles
with average diameter of about 10 nm were uniformly dispersed on the surface of WO3 . The EDS
results revealed that some particles were pure silver, and other particles were ascribed to AgCl,
which would be the direct evidence to prove the Ag/AgCl nanoparticles co-dispersed on the
surface of WO3 .
The photocatalytic activity of the Ag-AgCl/WO3 prepared by different photoreduction dura-
tion was evaluated in the degradation of 4-CP under the same condition as that used for Ag/AgCl
294 W.-L. Dai et al.

Figure 16.10. TEM images (A, B) and EDS (C) of 30Ag–AgCl/WO3 photocatalyst (Ma et al., 2012).

Figure 16.11. Photodegradation of 4-CP in aqueous suspension under visible light irradiation over different
photocatalysts (a) blank; (b) WO3 ; (c) 10Ag–AgCl/WO3 ; (d) 20Ag–AgCl/WO3 ; (e) 30Ag–
AgCl/WO3 ; (f) 40Ag–AgCl/WO3 (Ma et al., 2012).

core-shell sphere (as can be seen in Fig. 16.11). The reaction showed limited photocatalytic
activity for the degradation of 4-CP with WO3 alone, whereas the Ag-AgCl/WO3 catalysts
exhibited much higher photodegradation activity than WO3 . The photodegradation activity order
is 30Ag-AgCl/WO3 > 20Ag-AgCl/WO3 > 40Ag-AgCl/WO3 > 10Ag-AgCl/WO3 . Again, it was
demonstrated that the ratio of Ag and AgCl was a significant factor influencing the photodegrada-
tion rate. While the amount of metallic silver was very scarce, the generation of the electron-hole
pairs on the metallic Ag may be decreased due to the SPR effect on the metallic Ag, and the active
species was also cut down, which could reduce the photocatalytic activity. However, when the
Ag/AgCl composite material 295

Figure 16.12. Recycling test of 30 Ag-AgCl/WO3 under visible light irradiation (Ma et al., 2012).

Figure 16.13. TEM image of the as-prepared Ag/AgCl@TiO2 -20 photocatalyst (Guo et al., 2012).

amount of Ag was too high, some of the Ag nanoparticles became the active sites promoting the
recombination of the photoelectrons and the holes, which also reduced the photocatalytic activity.
Furthermore, the plasmonic photocatalystAg-AgCl/WO3 is investigated by the recycling exper-
iments. As presented in Figure 16.12, after six runs for the photodegradation of 4-CP, the sample
did not show significant loss of photocatalytic activity, which indicated the catalyst remained
stable during the photocatalytic reaction. In addition, 30Ag-AgCl/WO3 -hollow sphere photocat-
alyst exhibited higher photocatalytic activity than the Ag-AgCl/WO3 -nanoparticle, suggesting
that the structure, hollow sphere, definitely benefited the photocatalytic activity. All the results
above revealed that the as-prepared Ag-AgCl/WO3 is an active and stable visible-light-driven
plasmonic catalyst, which may serve as a promising candidate for practical use in the degradation
of pollutants.

16.5.4 Ag-AgCl@TiO2
Figure 16.13 showed the TEM images of Ag-AgCl@TiO2 -20. TEM images of other samples were
also taken, but no obvious changes were observed. Thus, only a single typical sample is shown
in this work. From Figure 16.13, it is shown that Ag0 NPs were well dispersed on the surface of
Ag-AgCl@TiO2 -20 with diameters ranging from 4 to 5 nm, implying that the synthesis method
reported in the present work was efficient for the preparation of nano-particles with unique particle
size distribution. The interlayer spacing d = 0.23 nm corresponded to the (111) plane of cubic
296 W.-L. Dai et al.

Figure 16.14. Photocatalytic activity of Ag-AgCl@TiO2 for photodegradation 4-CP under visible light
irradiation (a) blank; (b) N-TiO2 ; (c) Ag/TiO2 ; (d) Ag-AgCl@TiO2 -5; (e) Ag-AgCl@TiO2 -
10; (f) Ag-AgCl@TiO2 -20; (g) Ag-AgCl@TiO2 -30 (Guo et al., 2012).

Figure 16.15. Photocatalytic activity of Ag-AgCl@TiO2 for photo-reduction of Cr(VI) under visible light
irradiation (a) blank; (b) N-TiO2 ; (c) Ag/TiO2 ; (d) Ag-AgCl@TiO2 -5; (e) Ag-AgCl@TiO2 -
10; (f) Ag-AgCl@TiO2 -20; (g) Ag-AgCl@TiO2 -30 (Guo et al., 2012).

Ag, as shown in Figure 16.13B. Due to the low contrast of AgCl relative to TiO2 NPs in the TEM
images, it cannot be well distinguished from other NPs in the present work.
To evaluate the plasmon-induced photocatalytic activity of Ag-AgCl@TiO2 , the photodegra-
dation of 4-CP and photoreduction of Cr(VI) were also carried out under visible light irradiation
(Figs. 16.14 and 16.15) (Guo et al., 2012). Commercial P25 was tested as a reference prior to the
activity test of the as-prepared samples, and it was found that P25 showed almost no photodegra-
dation activity of 4-CP and Cr(VI) under identical conditions. Photocatalytic activity evaluations
were also performed with Ag/TiO2 and N-TiO2 suspensions under identical conditions for com-
parison. Ag-AgCl@TiO2 -20 exhibited the best photocatalytic activity among the three kinds of
Ag-AgCl@TiO2 samples. For the photodegradation of 4-CP, given that the degradation follows
a pseudo first order reaction pattern, the degradation rate of 4-CP over Ag-AgCl@TiO2 -5, Ag-
AgCl@TiO2 -10, Ag-AgCl@TiO2 -20, and Ag-AgCl@TiO2 -30 were estimated to be 0.085, 0.095,
0.114, and 0.057 min−1 , respectively, while the degradation rate over Ag/TiO2 and N-TiO2 were
only 0.017 and 0.0048 min−1 . Regarding the photo-reduction of Cr(VI), the results shown in
Figure 16.15 show the same trend as those from 4-CP. With increasing the duration of photo-
reduction, the Ag NP content increased correspondingly, changing with the ratios of Ag0 /Ag+ .
There should be an optimum ratio related to the most pronounced photoactivity. For a fair com-
parison of all the Ag-AgCl/TiO2 samples, the ratio of the Ag-AgCl/TiO2 -20 sample could best
Ag/AgCl composite material 297

approach the optimum one. That is the reason why the Ag-AgCl/TiO2 -30 sample shows lower
efficiency in photocatalytic activity.
In accordance with the previous report, during the photo-reduction of Cr(VI), 0.2 mM EDTA
was employed as sacrificial electron donator. The reduction became very slow without EDTA,
which has been verified by our experiment. According to the previous report, the conjugated
oxidation reaction of metal ion reduction is the electrochemical oxidation of water in the absence
of other organic species (Navío et al., 1998). This is a kinetically slow four-electron process and
we expected that the addition of sacrificial electron donors may accelerate the photocatalytic
reduction of metal ions. As a result, photocatalytic reduction in metal/organic photocatalyst
systems must be more efficient than in isolated metal photocatalyst systems. When metal and
organic photocatalysts coexisted in a system, organic species accept holes from the valence bands
directly or indirectly, thus suppressing the electron-hole recombination or increasing reduction
efficiency. The result of this work confirmed the expectation. In this way, the addition of EDTA
enhanced the Cr(VI) reduction very effectively. EDTA is regarded as a strong chelating agent,
which may form a stable complex with Cr(VI) and adsorb onto the surface of Ag-AgCl@TiO2
rather than conduct the photocatalytic reduction. To get rid of this possibility, physical adsorption
of Cr(VI) was conducted in a solution containing 0.2 mM EDTA, the percentage of Cr(VI) removed
was observed by dark adsorption and found to be only 12.3% within 1 h.
The stability of the plasmon photocatalyst Ag/AgCl@TiO2 , which is as important as its photo-
catalytic activity, was here investigated through recycling experiments. The result showed that the
sample did not show any significant loss of photocatalytic activity after five cycles of photodegra-
dation of 4-CP, indicating that the catalyst can keep stable during the photocatalytic reaction. All
of the results above indicated that the as-prepared Ag/AgCl@TiO2 is an active and stable visible-
light-driven plasmonic catalyst, which could serve as a promising candidate for practical use in
the degradation of hazardous pollutants in water.

16.5.5 Ag-AgI/Fe3 O4 @SiO2


Figure 16.16 shows the TEM images of different stages of Ag-AgI/Fe3 O4 @SiO2 composite
material in the preparation process. The diameter of the Fe3 O4 particles ranged from 150 to
300 nm and possessed regular sphere shape. After coated with SiO2 layer, core-shell Fe3 O4 @SiO2
microspheres with a thin silica layer of 20 nm in thickness were obtained. It is obviously
observed from Figures 16.16b and 16.16c that Ag0 NPs were well dispersed on the surface of
AgI/Fe3 O4 @SiO2 with diameter ranged from 10–20 nm in 90Ag-AgI/Fe3 O4 @SiO2 photocatalyst.
EDS spectra further confirmed the composite structure. Since the contrast of AgI compared to
Fe3 O4 @SiO2 NPs in TEM images is low, it cannot be distinguished from other NPs in the present
study.
To evaluate the photocatalytic activity of Ag-AgI/Fe3 O4 @SiO2 , the photodegradation of
RhB and 4-CP was carried out in aqueous dispersions under visible light irradiation (Guo
et al., 2011). RhB degradations were also performed in AgI/Fe3 O4 @SiO2 and Ag/Fe3 O4 @SiO2
suspensions under the same conditions for comparison. While irradiated at wavelengths of
λ > 420 nm, neither visible light nor Ag/Fe3 O4 @SiO2 showed any activity for the RhB degrada-
tion. 90Ag-AgI/Fe3 O4 @SiO2 exhibited the highest photocatalytic activity among the three kinds
of Ag-AgI/Fe3 O4 @SiO2 , which included 30Ag-AgI/Fe3 O4 @SiO2 , 90Ag-AgI/Fe3 O4 @SiO2 and
150 Ag-AgI/Fe3 O4 @SiO2 . The photodegradation rate over 90Ag-AgI/Fe3 O4 @SiO2 was about
18.4 times higher than that over AgI/Fe3 O4 @SiO2 and 10.9 times higher than that over N-TiO2 .
As for photodegradation of 4-CP, the results exhibited similar trend, except that the photocatalytic
activity of AgI/Fe3 O4 @SiO2 was slightly higher than that of N-TiO2 after reacted for 180 min.
This exception can be attributed to the appearance of more Ag0 nanoparticles on the surface of
AgI/Fe3 O4 @SiO2 , as a result of the long time of light irradiation. The photodegradation rate of
4-CP over 90Ag-AgI/Fe3 O4 @SiO2 was about 4.1 times higher than that over N-TiO2 . The above
results implied that the enhanced photocatalytic activity of Ag-AgI/Fe3 O4 @SiO2 can be ascribed
to the plasmon resonance of Ag species rather than the result of electron trapping by Ag NPs
298 W.-L. Dai et al.

(a) (b)

100 nm 100 nm

(c) Ag (d)
SiO2

Fe3O4 Intensity (n.u.) 0

C Fe
Si
Cu
Cu
Ag l
Ag Ag l l l Fe Cu

20 nm 0 1 2 3 4 5 6 7 8 9 10
Energy/keV

Figure 16.16. TEM images of (a) Fe3 O4 @SiO2 nanoparticles, (b) and (c) 90 Ag-AgI/Fe3 O4 @SiO2 and (d)
EDS of Ag-AgI/Fe3 O4 @SiO2 (Guo et al., 2011).

Figure 16.17. Pictures for progressive separation of Ag-AgI/Fe3 O4 @SiO2 from the suspension: (a) with-
out magnetic field presence, (b) upon application of a magnet for 0.5 min, and (c) upon
application of the magnet for 1 min (Guo et al., 2011).

and enhanced electron-hole separation. Though the Ag NPs can induce plasmon resonance, the
Ag/Fe3 O4 @SiO2 did not show obvious activity because no electron transfer occurred from Ag NPs
to the supported Fe3 O4 @SiO2 , which implied the AgI also play an important role in the enhanced
photocatalytic activity of Ag-AgI/Fe3 O4 @SiO2 . For Ag-AgI/Fe3 O4 @SiO2 , different photoreduc-
tion time in the preparation process can result in different Ag amount or NPs size on the surface
of AgI/Fe3 O4 @SiO2 . Due to the variation between the three kinds of Ag-AgI/Fe3 O4 @SiO2 , there
should be an optimal Ag loading in Ag-AgI/Fe3 O4 @SiO2 . The excess Ag0 would result in its
lower dispersion, then causing lower activity. Bearing in mind the rule that Ag0 can be dissolved
in HNO3 solution and AgI can be dissolved in concentrated KI solution, the ratio of Ag0 to AgI has
been determined by atomic absorption spectroscopy method. Consequently, the ratio of 30Ag-
AgI/Fe3 O4 @SiO2 , 90Ag-AgI/Fe3 O4 @SiO2 and 150Ag-AgI/Fe3 O4 @SiO2 were 3.4%, 6.7% and
8.1%, respectively.
Ag/AgCl composite material 299

Figure 16.17 presented the photos for magnetically induced separation of Ag-AgI/Fe3 O4 @SiO2
by exploiting their magnetic property. 0.01 g of sample was dispersed in 5 mL deionized water
and the suspension was sonicated for 10 min. The obtained suspension appeared to be dark yellow
in color (Fig. 16.17a). While placing the vial near an external magnet for 0.5 min, the progressive
separation of catalyst NPs was observed, while the catalyst was attracted toward the wall of the vial,
which was closer to the magnet (Fig. 16.17b). When the time duration was extended to 1 min, the
suspension became more clear (Fig. 16.17c). In addition, the catalyst powder can be re-dispersed
again simply by shaking after removing the magnet. Therefore, this kind of magnetic photocatalyst
NPs has the potential to be easily recovered after liquid phase photocatalytic reaction, which could
greatly facilitate the practical running of an industrial pollutant cleanup.

16.6 CONCLUSIONS

In summary, the plasmonic photocatalysts Ag/AgCl core-shell sphere, Ag/AgCl@Cotton-fabric,


Ag/AgCl@WO3 hollow sphere, Ag/AgCl@TiO2 , Ag/AgI/ Fe3 O4 @SiO2 , were successfully syn-
thesized by facile and effective method, exhibited highly efficient photocatalytic activity for the
degradation of RhB, 4-CP and / or Cr(VI) ion. The TEM and SEM images both confirmed the
core-shell structure of these catalysts. The ratio of Ag+ :Ag was found to influence the stability and
the photocatalytic activity of the as-prepared materials. All of the results above suggested that the
plasmon photocatalysts were active and stable as visible-light-driven plasmonic materials, which
could serve as promising candidates for practical use in the degradation of hazardous pollutants
in water.

ACKNOWLEDGEMENTS

This work was financially supported by the Major State Basic Resource Development Program
(Grant No. 2012CB224804), NNSFC (Project 20973042, 21173052), and the Natural Science
Foundation of Shanghai Science and Technology Committee (08DZ2270500).

REFERENCES

Abe, R., Sayama, K. &Arakawa, K.: Efficient hydrogen evolution from aqueous mixture of I− and acetonitrile
using a merocyanine dye-sensitized Pt/TiO2 photocatalyst under visible light irradiation. Chem. Phys.
Lett. 362:5–6 (2002), pp. 441–444.
Awazu, K., Fujimaki, M., Rockstuhl, C., Tominaga, J., Murakami, H., Ohki, Y., Yoshida, N. & Watanabe, T.:
A plasmonic photocatalyst consisting of silver nanoparticles embedded in titanium dioxide. J. Am. Chem.
Soc. 130:5 (2008), pp. 1676–1680.
Bessekhouad, Y., Chaoui, N., Trzpit, M., Ghazzal, N., Robert, D. & Weber, J.: UV-vis versus visible degra-
dation of Acid Orange II in a coupled CdS/TiO2 semiconductors suspension. J. Photochem. Photobiol.
A: Chem. 183:1–2 (2006), pp. 218–224.
Bi, Y.P. & Ye, J.H.: In situ oxidation synthesis of Ag/AgCl core-shell nanowires and their photocatalytic
properties. Chem. Commun. (2009), pp. 6551–6553.
Chen, D. & Ye, J.: Hierarchical WO3 hollow shells: Dendrite, sphere, dumbbell, and their photocatalytic
properties. Adv. Funct. Mater. 18:13 (2008), pp. 1922–1928.
Chen, H.Y., Nambu, A., Wen, W., Graciani, J., Zhong, Z., Hanson, J.C., Fujita, E. & Rodriguez, J.A.:
Reaction of NH3 with titania: N-doping of the oxide and TiN formation. J. Phys. Chem. C 111:3 (2006),
pp. 1366–1372.
Chen, Y.S., Crittenden, J.C., Hackney, S., Sutter, L. & Hand, D.W.: Preparation of a novel TiO2 -based p-n
Junction nanotube photocatalyst. Environ. Sci. Technol. 39:5 (2005), pp. 1201–1208.
Davis, P. & Green, D.L.: Photocatalytic oxidation of cadmium-EDTA with titanium dioxide. Environ. Sci.
Technol. 33:4 (1999), pp. 609–617.
300 W.-L. Dai et al.

Du, J., Lai, X.Y., Yang, N.L., Zhai, J., Kisailus, D., Su, F.B., Wang, D. & Jiang, L.: Hierarchically
ordered macro-mesoporous TiO2 −graphene composite films: Improved mass transfer, reduced charge
recombination, and their enhanced photocatalytic activities. ACS Nano 5:1 (2011), pp. 590–596.
Fang, X.M., Zhang, Z.G., Chen, Q.L., Ji, H.B. & Gao, X.N.: Dependence of nitrogen doping on TiO2
precursor annealed under NH3 flow. J. Solid State Electrochem. 180:4 (2007), pp. 1325–1332.
Fox, M.A. & Dulay, M.T.: Heterogeneous photocatalysis. Chem. Rev. 93:1 (1993), pp. 341–357.
Fujishima, A. & Honda, K.: Electrochemical photolysis of water at a semiconductor electrode. Nature
238:5383 (1972), pp. 37–38.
Fujishima, A, Nagahara, L.A., Yoshiki, H., Ajito, K. & Hashimoto, K.: Thin semiconductor-films-photo
effects and new applications. Electrochim. Acta 39:8–9 (1994), pp. 1229–1236.
Guo, J.F., Ma, B.W., Yin, A.Y., Fan, K.N. & Dai, W.L.: Photodegradation of rhodamine B and 4-
chlorophenol using plasmonic photocatalyst of Ag-AgI/Fe3 O4 @SiO2 magnetic nanoparticle under visible
light irradiation. Appl. Catal. B 101:3–4 (2011), pp. 580–586.
Guo, J.F., Ma, B.W., Yin, A.Y., Fan, K.N. & Dai, W.L.: Highly stable and efficient Ag/AgCl@TiO2 photo-
catalyst: Preparation, characterization, and application in the treatment of aqueous hazardous pollutants.
J. Hazard. Mater. 211–212 (2012), pp. 77–82.
Haick, H. & Paz, Y.: Long-range effects of noble metals on the photocatalytic properties of titanium dioxide.
J. Phys. Chem. B 107:10 (2003), pp. 2319–2326.
Hoffmann, M.R., Martin, S.T., Choi, W.Y. & Bahnemann, D.W.: Environmental applications of semiconduc-
tor photocatalysis. Chem. Rev. 95:1 (1995), pp. 69–96.
Hu, C., Peng, T.W., Hu, X.X., Nie, Y.L., Zhou, X.F., Qu, J.H. & He, H.: Plasmon-induced photodegradation
of toxic pollutants with Ag−AgI/Al2 O3 under visible-light irradiation. J. Am. Chem. Soc. 132:2 (2010),
pp. 857–862.
Hurum, D.C., Gray, K.A., Rajh T. & Thurnauer, M.C.: Recombination pathways in the Degussa P25
formulation of TiO2 : Surface versus lattice mechanisms. J. Phys. Chem. B 109:2 (2005), pp. 977–980.
Kakuta, N., Goto, N., Ohkita, H. & Mizushima, T.: Silver bromide as a photocatalyst for hydrogen generation
from CH3 OH/H2 O solution. J. Phys. Chem. B 103:29 (1999), pp. 5917–5919.
Lee, M.K. & Shih, T.H.: High photocatalytic activity of nanoscaled heterojunction of ZnS grown on fluorine
and nitrogen co–doped TiO2 . J. Electrochem. Soc. 154:4 (2007), pp. 49–51.
Li, F., Liu, X.Q., Yuan, Y.L., Wu, J.F. & Li, Z.: Controllable synthesis, characterization and optical properties
of Ag@AgCl coaxial core-shell nanocables. Cryst. Res. Technol. 45:11 (2010), pp. 1189–1193.
Lu, Z.X., Zhou, L., Zhang, Z.L., Shi, W.L., Xie, Z.X., Xie, H.Y., Pang, D.W. & Shen, P.: Cell damage induced
by photocatalysis of TiO2 thin films. Langmuir 19:21 (2003), pp. 8765–8768.
Ma, B.W., Guo, J.F., Zou, L.Y., Dai, W.L. & Fan, K.N.: Ag/AgCl@Cotton-fabric: A highly stable and
easy-recycling plasmonic photocatalyst under visible light irradiation. Chin. J. Chem. 29:4 (2011),
pp. 857–859.
Ma, B.W., Guo, J.F., Dai, W.L. & Fan, K.N.: Ag-AgCl/WO3 hollow sphere with flower-like structure and
superior visible photocatalytic activity. Appl. Catal. B: Environ. 123–124 (2012), pp. 193–199.
Ma, B.W., Guo, J.F., Dai, W.L. & Fan, K.N.: Highly stable and efficient Ag/AgCl core-shell sphere: Con-
trollable synthesis, characterization, and photocatalytic application. Appl. Catal. B: Environ. 130–131
(2013), pp. 257–263.
Mrowetz, M., Balcerski, W., Colussi, A.J. & Hofmann, M.R.: Oxidative power of nitrogen-doped TiO2
photocatalysts under visible illumination. J. Phys. Chem. B 108:45 (2004), pp. 17269–17273.
Navío, J.A., Colón, G., Trillas, M., Peral, J., Domenech, X., Testa, J.J., Padrón, J., Rodríguez, D. & Litter,
M.I.: Heterogeneous photocatalytic reactions of nitrite oxidation and Cr (VI) reduction on iron-doped
titania prepared by the wet impregnation method, Appl. Catal. B 16:2 (1998), pp. 187–196.
O’Regan, B. & Graetzel, M.: A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2
films. Nature 353:6346 (1991), pp. 737–740.
Sato, S., Nakamura, R. & Abe, S.: Visible-light sensitization of TiO2 photocatalysts by wet-method N doping.
Appl. Catal. A: Gen. 284:1–2 (2005), pp. 131–137.
Soni, S.S., Henderson, M.J., Bardeau, J. F. & Gibaud, A.: Visible-light photocatalysis in titania-based
mesoporous thin films. Adv. Mater. 20:8 (2008), pp. 1493–1498.
Tang, J.W., Zou, Z.G. & Ye, J.H.: Efficient photocatalytic decomposition of organic contaminants over
CaBi2 O4 under visible-light irradiation. Angew. Chem. Int. Ed. 43:34 (2004), pp. 4463–4466.
Wang, C., Zhao, J.C., Wang, X.M., Mai, B.X., Sheng, G.Y., Peng, P. & Fu, J.M.: Preparation, characterization
and photocatalytic activity of nano-sized ZnO/SnO2 coupled photocatalysts. Appl. Catal. B: Environ. 39:3
(2002), pp. 269–279.
Ag/AgCl composite material 301

Wang, P., Huang, B.B., Qin, X.Y., Zhang, X.Y., Dai, Y., Wei, J.Y. & Whangbo, M.H.: Ag@AgCl: A
highly efficient and stable photocatalyst active under visible light. Angew. Chem. Int. Ed. 47:41 (2008),
pp. 7931–7933.
Wang, P., Huang, B., Zhang, X.Y., Qin, X.Y., Jin, H., Dai, Y., Wang, Z.Y., Wei, J.Y., Zhan, J., Wang, S.Y.,
Wang, J.P. & Whangho, M.H.: Highly efficient visible-light plasmonic photocatalyst Ag@AgBr. Chem.
Euro. J. 15:8 (2009), pp. 1821–1824.
Wang, S., Yi, L.X., Halpert, J.E., Lai, X.Y., Liu, Y.Y., Cao, H.B., Yu, R.B., Wang D. & Li, Y.L.: A novel and
highly efficient photocatalyst based on P25-graphdiyne nanocomposite. Small 8:2 (2012), pp. 265–271.
Wittbrodt, P.R. & Palmer, C.D.: Reduction of Cr(VI) in the presence of excess soil fulvic acid. Environ. Sci.
Technol. 29:1 (1995), pp. 255–263.
Yan, X.L., Ohno, T., Nishilima, K., Abe, R. & Ohtani, B.: Is methylene blue an appropriate substrate for
a photocatalytic activity test? A study with visible-light responsive titania. Chem. Phys. Lett. 429:4–6
(2006), pp. 606–610.
Yu, H.T., Quan, X., Chen, S. & Zhao, H.M.: TiO2 – multiwalled carbon nanotube heterojunction arrays and
their charge separation capability. J. Phys. Chem. C. 111:35 (2007), pp. 12,987–12,991.
This page intentionally left blank
CHAPTER 17

Highly photoactive Er3+ -TiO2 system by means of up-conversion


and electronic cooperative mechanism

Sergio Obregón & Gerardo Colón

17.1 INTRODUCTION

Heterogeneous photocatalysis technology is considered as a promising route to afford pollutant


degradation as well as hydrogen production processes (Colón et al. 2007). Nowadays, the research
activity in this field is focused in the development of novel alternative materials to traditional
TiO2 capable to use of sunlight as the green energy source (Kubacka et al. 2012). In this sense, the
utilization of solar light as efficient as possible has been largely pursued. For this scope, different
strategies have been traditionally followed, based in all cases in the improvement of visible photon
absorption. Among these, modification of TiO2 by anion-cation codoping or coupling TiO2 with
other semiconductors with low band gaps have proved to be viable ways to allow the extension
of light absorption edge (Li et al. 2007, Kubacka et al. 2009, Hidalgo et al. 2009 Colón et al.
2010). An alternative option consists on the creation of new single phase visible active catalysts
which would overcome the drawbacks of doping (Zhang et al. 2005, Zhang et al. 2007, Song
et al. 2011, Naya et al. 2011, Obregón et al. 2012).
Other challenging composite configuration would consist on the assembly of the photocatalytic
material with a luminescence material (Zhang et al. 2011, Zhou et al. 2011, Li et al. 2011). Within
this configuration the improvement is reached by increasing the number of incoming radiation
photons absorbed by the photocatalyst. This approach appears as a completely new alternative for
enhancing the efficiency of the photocatalytic process by a wise light handling. Thus, the lumines-
cent material could handle and transform the not absorbed photons into appropriate energy ones
(Kenyon, 2002). Considering an up-converting luminescent material, it would absorb low energy
radiation (i.e. on the NIR or visible range) and emit higher energy radiation (i.e. in the visible
and/or UV). In this case, two/three low-energy photons are ‘added up’ to give one higher-energy
photon. The applications of the up-conversion process by phosphor-like systems (e.g. NIR or visi-
ble to UV) would optimize the photocatalytic performance of traditional UV active photocatalysts.
Recently, Qin et al. reported the combination of YF3 :Yb3+ ,Tm3+ luminescent nanoparticles with
TiO2 which exhibits near-infrared photoactivity (Qin et al. 2010). Similarly, Li et al. proposed
Y2 O3 :Yb3+ ,Tm3+ ‘phosphor’ as NIR to UV up-conversion promoter (Li et al. 2011).
Among various up-converting nanomaterials, Er3+ constitute an interesting option for this
purpose which could be excited by visible-light or NIR, showing luminescence in the visible and
ultraviolet (Wang et al. 2006). Therefore, the assembly of TiO2 with such up-converting doping
cation should in principle provide extra UV photons and increase the photocatalytic activity. In
the present paper, we propose the Er3+ doping into TiO2 as host matrix. This way we avoid the
formation of composite system which could lead to unfavorable charge handling process.

17.2 EXPERIMENTAL SECTION

17.2.1 Synthesis of photocatalysts


TiO2 sample was obtained by means of a hydrothermal method elsewhere described. In brief,
a TiO2 colloidal solution was obtained by adding certain amount of Ti4+ -isopropanol solution
303
304 S. Obregón & G. Colón

(38.4 mL of Ti4+ isopropoxide +38.2 mL of isopropanol) to 400 mL of distilled water at pH = 2


achieved by means of acetic acid. After isopropoxide addition, a white precipitate is obtained
that upon stirring at room temperature for one week evolves to a milky homogeneous solution. A
certain amount of triethylamine (TEA) was then added drop wise to the Ti-solution aliquot till the
pH value was 9. Afterwards, the obtained white precipitate suspension was then placed in a Teflon
recipient inside of stainless steel autoclave reactor. The hydrothermal treatment was performed
at 140◦ C, 20 hours. The as obtained precipitate was then filtered, repeatedly washed and dried
overnight at 120◦ C. Then TiO2 powder was submitted to a further calcination treatment at 300◦ C
for 2 hours.
Erbium doped TiO2 were obtained based on the above described methods for the preparation
of single pristine photocatalysts. Thus, to the TiO2 solution the appropriate amount of Er(NO3 )3
was added in order to accomplish the desired Er3+ -TiO2 relationship (0.5, 1, 2, 3 and 4 at%).
Then, a similar procedure is followed to achieve to the final photocatalyst.

17.2.2 Materials characterization


BET surface area and porosity measurements were carried out by N2 adsorption at 77 K using a
Micromeritics 2010 instrument.
X-ray diffraction (XRD) patterns were obtained using a Siemens D-501 diffractometer with
Ni filter and graphite monochromator. The X-ray source was Cu Kα radiation (0.15406 nm).
Rietveld analyses were performed by using XPert HighScore Plus software over selected samples.
The diffraction patterns were recorded from 2θ 10◦ to 120◦ with step of 0.017◦ and 400 s per step.
Crystallite sizes were obtained from Rietveld refinement.
Micro-Raman measurements were performed using a LabRAM Jobin Yvon spectrometer
equipped with a microscope. Laser radiation (λ = 532 nm) was used as excitation source at
5 mW. All measurements were recorded under the same conditions (2 s of integration time and 30
accumulations) using a 100× magnification objective and a 125 mm pinhole.
UV–vis spectra (Shimadzu, AV2101) were recorded in the diffuse reflectance mode (R) and
transformed to a magnitude proportional to the extinction coefficient (K) through the Kubelka–
Munk function, F(R∞). Samples were mixed with BaSO4 that does not absorb in the UV–vis
radiation range (white standard). Scans range was 240–800 nm. Assuming the catalysts as indirect
semiconductors it is possible to construct a plot (αhν)1/2 vs hν, also called Tauc plot (Tauc, 1970),
in order to get an accurate value of the band gap of the solids.
XPS data were recorded on 4 × 4 mm2 pellets, 0.5 mm thick, prepared by slightly pressing the
powered materials which were outgassed in the prechamber of the instrument at room temper-
ature up to a pressure <2·10−8 to remove chemisorbed water from their surfaces. The SPECS
spectrometer main chamber, working at a pressure <10−9 torr, was equipped with a PHOIBOS
100 multichannel hemispherical electron analyser with a dual X-ray source working with Mg Kα
(hν = 1253.6 eV) at 120 W, 20 mA using C 1s as energy reference (284.6 eV). Surface chemical
compositions were estimated from XP-spectra, by calculating the integral of each peak after sub-
traction of the “S-shaped” Shirley-type background using the appropriate experimental sensitivity
factors using UNIFIT 2012 software (Hesse, 2012).
The excitation and emission spectra of the catalysts were recorded at ambient temperature in
a Horiba Jobin-Yvon Fluorolog3 spectrofluorometer operating in the front face mode operating
with a 5 nm slit. The UC optical measurements were performed for powdered pressed samples
using a Jenoptik laser diode source at 980 nm.

17.2.3 Photocatalytic experimental details


Phenol oxidation reactions were performed using a batch reactor (250 mL) using an arc lamp
source (Oriel Instruments) equipped with an Hg-Xe lamp of 200 W. The intensity of the incident
UVA and visible light on the solution was measured with an HD2302 photometer (Delta OHM)
using LP 471 UVA and LP 471 RAD sensors (spectral responses 315–400 nm and 400–1050 nm
Highly photoactive Er3+ -TiO2 system 305

respectively). An oxygen flow was employed what produces a homogenous suspension of the
catalyst in the solution. Before each experiment, the catalysts (1 g L−1 ) were settled in suspension
with the reagent mixture for 15 min. The evolution of the initial phenol concentration (ca. 30 ppm)
was followed through the evolution of the characteristic 270 nm band using a centrifuged aliquot
ca. 2 mL of the suspension (microcentrifuge Minispin, Eppendorf). In order to distinguish the
different contributions of the different ranges of the lamp spectrum, UV-vis-NIR, UV-vis, visible-
NIR and visible photocatalytic experiments were performed by using a UV and IR cut-off filters
(λ < 420 nm and λ>800 nm respectively).
The photonic efficiency for the phenol decomposition reaction under UV and NIR has been
determined from the reaction rate and the flux of incoming photons (calculated for the irradi-
ation wavelengths of 360 nm and 980 nm) according to the following equation (Baumanis and
Bahnemann, 2008):
V · c
ζ=
J · A · t
where V = volume; c = change in concentration; J = flux of photons; A = illuminated area;
t = change in time.
The flux of photons was calculated by means of the following equation:
I ·λ
J=
NA · h · c

where I = light intensity (W/m2 ); λ = 360 or 980 nm; NA = Avogadro constant; h = Planck’s
constant; and c = speed of light.

17.3 RESULTS AND DISCUSSION

From the XRD patterns, it appears that the incorporation of the Er3+ cation does not produce any
significant change in the crystalline phase present. The hydrothermal preparation followed lead
to the anatase phase exclusively (PDF card 78-2486, corresponding to the I41 /amd space group);
no traces of rutile or brookite are observed (Figure 17.1). On the other hand, even for the highest
Er3+ content sample, no indication of Er2 O3 species is observable. The structural characterization
of Er3+ -doped TiO2 was performed by Rietveld analysis of the XRD diffraction patterns. From
this analysis we have obtained the lattice parameters of the tetragonal cell. It can be observed that
upon Er3+ incorporation a clear distortion of the lattice is achieved.
The incorporation of the Er3+ cation clearly affects the anatase lattice producing a certain
enlargement of the cell. Thus, the cell volume calculated from the lattice parameters shows a
slight increment with Er3+ content (Figure 17.1). This cell expansion is higher when Er3+ content
exceeds the 2 at%. It is clear that the anatase lattice tries to accommodate the doping cation
producing the observed expansion of the cell. In fact, by taking into account the ionic radius
value of Er3+ (89 pm) and Ti4+ (68 pm) it could be assumed that Er3+ could either replace some
Ti4+ sites or occupy the interstitial site of anatase structure. Similar effect was already reported
for Nd3+ (with ionic radius of 98 pm) doping of TiO2 (Li et al. 2005). Moreover, due to the
differences in the ionic radii between Er3+ and Nd3+ in our case the cell expansion was not so
pronounced. Other authors argued the Er3+ ions replacing the Ti4+ sites would have a stabilizing
effect on the Ti-O bond. Because of the more electropositive character of Er3+ , a strengthening
of the bonding between O2− and the less electropositive Ti4+ ions would take place (Reddy et al.
2001).
Table 17.1 summarizes the surface, structural data for Er3+ doped TiO2 samples. It can be noted
that the crystallite size for anatase phase is progressively reduced upon doping. This fact has been
already reported for lanthanide doped TiO2 . Thus, several authors reported that lanthanide doping
of TiO2 can decrease the crystallite size (Xu et al. 2002). The decrease in particle size has been
attributed to the presence of Ti-O-Ln bonds or lanthanide oxide particles at the interfaces in the
306 S. Obregón & G. Colón

Figure 17.1. XRD patterns of Er3+ -TiO2 catalysts obtained at pH = 9 after hydrothermal treatment at
140◦ C, 20 h. Inset plot shows the evolution of cell.

Table 17.1. Surface, structural and photocatalytic characterization for Er3+ doped TiO2 catalysts.

Er3+ content (%)


Cryst. size BET Av. Pore size Band gap
Samples (nm) (m2 g−1 ) (nm) (eV) EDX XPS O/Ti

TiO2 15 102 9 3.16 — — 1.90


Er-TiO2 0,5 at% 13 110 9 3.16 — 0.4 1.91
Er-TiO2 1 at% 14 110 11 3.18 1.2 0.9 1.90
Er-TiO2 2 at% 14 110 11 3.18 1.9 1.8 1.88
Er-TiO2 3 at% 13 112 12 3.18 2.7 2.8 1.85
Er-TiO2 4 at% 12 116 12 3.20 4.0 3.7 1.86

doped samples, which prevents anatase particles from adhering together and inhibits the growth
of crystal grains (Xu et al. 2002, Li et al. 2004)).
Regarding to the BET surface area, Er3+ -TiO2 samples exhibit relatively high values; being, in
any case, slightly higher than for undoped TiO2 . At the same time, the pore size presents a narrow
distribution with an average pore size around 9–12 nm in all cases with increasing average values
as Er3+ content rises (Table 17.1).
The evolution of the Er3+ -doped TiO2 has been also followed by Raman spectroscopy. From the
irreducible presentation of the optical modes it is possible to distinguish the three TiO2 polymorphs
by means of Raman spectroscopy (Colón et al. 2006). From the observation of the Raman spectra
shown in Figure 17.2, it is possible to infer that anatase is the only crystalline phase present in the
whole series. The shift in the position of the Raman lower energy Eg band and the broadening
in the FWHM of such band have been extensively discussed and related with different structural
defects of the material (Zhu et al. 2005, Spanier et al. 2001). In our case, the position of the
band at around 150 cm−1 suffers a small shift as Er3+ content increases (Figure 17.2). The close
correlation observed between the Raman Eg peak position and anatase cell volume (Figure 17.1)
Highly photoactive Er3+ -TiO2 system 307

Figure 17.2. Raman spectra of Er3+ -TiO2 catalysts obtained at pH = 9 after hydrothermal treatment at
140◦ C, 20 h. Inset: Detail of Eg1 band evolution with Er3+ content.

Figure 17.3. Diffuse reflectance spectra of Er3+ -doped TiO2 systems.

behavior throughout the sample series clearly indicates that the Raman observable is dominated by
structural effects resulting from Er incorporation to the anatase lattice, giving thus an additional
proof of the presence of the ion at the anatase structure.
In addition to the Raman bands above described, it can be observed other series corresponding
to the fluorescence of Er3+ upon excitation with Raman laser at 532 nm. The presence of these
bands denotes the luminescence behaviour of Er3+ ions on the TiO2 matrix.
Regarding to UV-vis absorption properties, the diffuse reflectance UV-vis spectra suggest that
Er3+ doping does not affect notably the absorption edge of the TiO2 (Figure 17.3). On this basis,
308 S. Obregón & G. Colón

Figure 17.4. a) Evolution of the reaction rates for phenol photodegradation for different Er3+ -TiO2 catalysts
under UV-vis-NIR, UV-vis and NIR irradiation; b) Photon efficiencies for phenol degradation
reaction under UV and NIR irradiation for different Er3+ -TiO2 catalysts.

we would state that the incorporation of Er3+ does not induce any appreciable alteration of the
band structure. Additionally to the TiO2 absorption spectrum, for doped systems different bands
appear in the range 400–700 nm which can be associated to the excitation of Er3+ species (Wang
et al. 2008). These absorption lines located at 489, 520, 653 and a small tail at 800 nm would
correspond to the transitions from the Er3+ ground state 4 I15/2 to the higher energy levels.
From XPS, the surface quantitative analysis provides a close correspondence with the values
obtained from EDX analysis (Table 17.1). This fact clearly denotes a homogeneous distribution of
Er3+ cation from the bulk to the surface. In other words, no Er3+ segregation might be expected by
taking into account these values. This result would support the proposed idea of the homogeneous
insertion of erbium into the anatase lattice (occupying substitutional or interstitial sites).
The photocatalytic performance of the Er-doped systems was studied for phenol degradation
upon irradiation with different spectral ranges of the lamp (Figure 17.4). By observing the con-
version plot under illumination with the full range of the lamp, it is clear that the incorporation
of Er3+ leads to a progressive increase in the photocatalytic activity. Thus, the best reaction rate
was attained for sample doped with 2 at% of Er3+ . In order to distinguish the contribution of each
region of the lamp spectrum, different experiments were carried out using UV and NIR filters.
Visible light runs (performed by combined use of UV and NIR filters) provide null photoactivity
for all samples. Furthermore, UV-vis photocatalytic experiments lead to a similar reaction rate
evolution with Er3+ content with respect to full range experiments. In this case, the reaction rates
appeared clearly diminished due to the cut-off effect of NIR filter along the whole spectrum of the
lamp which seems to affect in great manner to the UV active zone of the catalyst. From this result,
it can be pointed out that the erbium presence has a clear effect on the photocatalytic activity of
TiO2 under UV irradiation. Finally, we have suppressed the UV region of the lamp. This way we
obtained the here called vis-NIR photocatalytic runs (Figure 17.4). Since within the experiments
under vis light all samples appeared inactive, it is expected that the small photoactivity observed
under vis-NIR irradiation might be attributed to NIR photoactivation. By observing the conversion
plots of Er3+ -TiO2 for phenol under such conditions, it is clear that, though small, a progressive
increase in the reaction rate is taking place, having the maximum value for 2 at% of Er3+ content.
In Figure 17.4 we also show the calculated photon efficiencies for the reactions under UV-vis
and NIR. Although these calculations are only semiquantitative as they did not account for the
Highly photoactive Er3+ -TiO2 system 309

Scheme 17.1. Proposed photon and electronic mechanism upon UV excitation.

number of electrons involved in the photo-chemical reaction, the real calculation of the photon
efficiency taking into account this issue is, as demonstrated by Ohtani, rather complex (Ohtani
2008). In spite of this and as we are analyzing the same reaction for all experiments reported
here, the trends and relative values displayed are valid. Thus, from Figure 17.4 values it appears
that phenol degradation reaction proceeds more efficiently under NIR irradiation than under UV.
These results would clearly support the assumption made that Er3+ plays a significant role in the
overall photocatalytic mechanism. Thus, while for UV irradiation, samples with Er3+ content
higher than 1 at% gives similar photon efficiencies, under NIR, the ζ increase remarkably for
Er3+ loading higher than 1 at%. This fact would point out the participation of an up-conversion
process.
On the basis of the above photocatalytic results, a double cooperative mechanism would be
envisaged. On the one hand, under UV illumination, the presence of Er3+ in the anatase struc-
ture would affect in the electronic mechanism of the photocatalytic process initiated by TiO2
(Scheme 17.1). It has been reported that certain lanthanide ions incorporated in the TiO2 structure
could act as good scavengers to trap photoelectrons (Silva et al. 2009).
The incorporation of Er3+ into the anatase structure at Ti4+ positions would generate certain
amounts of oxygen vacancies in order to achieve the charge neutrality of the network. The quan-
titative determination of these oxygen vacancies is quite complex. The observation of the O/Ti
ratio calculated from XPS analysis is a semiquantitative approach which could give us an indica-
tion of the presence of such defective sites in the structure (Table 17.1). By observing the O/Ti
calculated from XPS results, it is clear that upon Er3+ incorporation the decreasing trend of the
O/Ti would indicate the occurrence of oxygen vacancies at the structure. Thus, it may be assumed
that upon irradiation the photogenerated electrons in the TiO2 could be trapped by such interband
defect states lying below the conduction band edge of TiO2 . In this sense, Bettinelli et al. also
reported that for lanthanide doped TiO2 there is an important role of the traps for the transit of the
electrons in the doped systems (Bettinelli et al. 2006). These authors argued that a small fraction
of trapping/detrapping process would be the responsible of the diffusion coefficient lowering.
On the other hand, upon NIR irradiation, the presence of Er3+ cation seems to promote an
up-conversion process, pumping photons in the UV range into the TiO2 structure (Scheme 17.2).
It has been reported that Er3+ shows a small UV photoluminescence emission at around 390 nm
after excitation at 980 nm (Wang et al. 2008). This up-conversion process would involve a
sequential three photons absorption (4 I15/2 →4 I11/2 , 4 I11/2 →4 F7/2 and 4 S3/2 →2 G7/2 ). Then, by
a multiphonon relaxation the 2 G7/2 excited state decays to 2 G11/2 and 2H9/2 lower states. The
photoluminescence emission in the UV /violet range is then produced by the 2 G11/2 →4 I15/2 tran-
sition, giving a small emission at around 390 nm. Thus, it can be assumed that the improvement
of the photoefficiency might be related to the increasing the number of available photons with
the appropriate energy.
310 S. Obregón & G. Colón

Scheme 17.2. Proposed photon and electronic mechanism upon NIR excitation.

Figure 17.5. a) Er3+ luminescent spectra under λex = 980 nm excitation, and b) PL emission spectra (λex =
290 nm) of TiO2 and Er3+ -TiO2 (2 at%) photocatalysts.

In order to support the proposed doubled mechanism we have studied the photoluminesce prop-
erties of Er3+ -TiO2 systems. Thus, by excitation of samples with 980 nm light the photoluminesce
(PL) spectra in the UV region show two small emission bands located at ca. 390 and 415 nm (Fig-
ure 17.5.a). This evidence would confirm the up-conversion process proposed in which Er3+
would convert the NIR photons in high energy UV photons. On this basis, these up-converted
photons would be absorbed by TiO2 enhancing the photoactivity for phenol degradation.
Photocatalytic activity is closely related to the efficiency of the photogenerated electron–hole
separation and the transfer from the inner regions to the outer surfaces. Thus, PL emission spectra
have been also widely used to investigate the efficiency of charge carrier trapping, migration,
Highly photoactive Er3+ -TiO2 system 311

transfer, and to understand the fate of photogenerated electrons and holes in the semiconductor
(Li et al. 2005, Mercado et al. 2011). In this sense, by observing the PL spectra obtained after
excitation with 290 nm light, TiO2 and Er3+ -TiO2 systems exhibited emission in the range of 350–
550 nm (Figure 17.5.b). This broad luminescence band could be attributed to defect levels near
the valence band, oxygen vacancy with two trapped electrons (i.e. F center), oxygen vacancies
with one trapped electron (Mercado et al. 2011). From these plots it can be noticed that Er3+ -TiO2
exhibits a drastic, general decrease in the PL intensity, which clearly indicates that the presence of
Er3+ species induces a hard suppression of the recombination process of photogenerated electrons
and holes.

17.4 CONCLUSIONS

We have obtained a highly active TiO2 system by doping with Er3+ . The hydrothermal synthesis
leads to homogeneous TiO2 nanoparticles of around 10 nm size. From the structural and surface
analysis the incorporation of Er3+ can be stated. The presence of Er3+ species produces a slight
distortion of the anatase lattice. From the combined use of photocatalytic and luminescence
experiments we suggest a dual-type mechanism which could explain the improved photoactivities
in doped systems. On one hand, Er3+ improved the electronic charge separation process, enhancing
the UV -photoassisted process. This mechanism is consistent with a classical doping process using
TiO2 as host material. However, it can be evidenced an additional, small contribution of NIR
photons in the overall mechanism due an energy transfer process from erbium ions to TiO2 . Thus,
better photon efficiencies have been found for the reaction under NIR irradiation with respect
to reaction under UV conditions. As a consequence of these two processes, strongly improved
photoactivities have been found for liquid phase phenol degradation reactions, indicating the Er
positive influence in anatase TiO2 chemical activity.

ACKNOWLEDGEMENTS

The financial support by projects P09-FQM-4570 and ENE2011-24412 is fully acknowledged.


S. Obregón thanks CSIC for the concession of a JAE-Pre grant. The authors thank H. Miguez and
M. Calvo (ICMSE, Seville Spain) for photoluminescence measurements and for their valuable
comments.

REFERENCES

Bettinelli, M., Speghini, A., Falcomer, M., Daldosso, M., Dallacasa, V. & Romanò, L.: Photocatalytic, Spec-
troscopic and Transport Properties of Lanthanide-doped TiO2 Nanocrystals. J. Phys. Condens. Matter.
18, (2006), S2149.
Colón, G., Hidalgo, M.C., Munuera, G., Ferino, I., Cutrufello, M.G. & Navío, J.A.: Structural and Surface
Approach to the Enhanced Photocatalytic Activity of Sulfated TiO2 Photocatalyst. Appl. Catal. B: Environ.
63, (2006), 45–59.
Colón, G., Belver, C. & Fernández-García, M.,“Nanostructured Oxides in Photocatalysis”, in: Synthesis,
Properties and Application of Oxide Nanomaterials. Eds. M. Fernández-García, J.A. Rodríguez. Wiley,
USA, 2007. (ISBN: 978-0-471-72405-6).
Hesse, R.: Unifit for Windows: Spectrum Processing, Analysis and Presentation Software for Photoelectron
Spectra, Version 2012, Leipzig. http://www.unifit-software.de.
Kenyon, A.J.: Recent developments in rare-earth doped materials for optoelectronics. Prog. Quantum
Electron. 26, (2002), 225–284.
Kubacka, A., Fernández-García, M. & Colón, G.: Advanced Nanoarchitectures for Solar Photocatalytic
Applications. Chem. Rev. 112, (2012), 1555–1614.
Li, F.B., Li, X.Z. & Hou, M.F.: Photocatalytic Degradation of 2-mercaptobenzothiazole in Aqueous La3+ –
TiO2 Suspension for Odor Control. Appl Catal B: Environ. 48, (2004), 185–194.
312 S. Obregón & G. Colón

Li, W., Frenkel, A.I., Woicik, J.C., Ni, C. & Shah, S.I.: Dopant Location Identification in Nd3+ -doped TiO2
Nanoparticles. Phys. Rev. B 72, (2005), 155315.
Li, D., Haneda, H., Hishita S. & Ohashi, N.: Visible-Light-Driven N-F-Codoped TiO2 Photocatalysts. 2.
Optical Characterization, Photocatalysis, and Potential Application to Air Purification. Chem. Mater.,
2005, 17, 2596–2602.
Li, T., Liu, S., Zhang, H., Wang, E., Song, L. & Wang, P.: Ultraviolet Upconversion Luminescence in
Y2 O3 :Yb3+ ,Tm3+ Nanocrystals and its Application in Photocatalysis. J. Mater. Sci. 46, (2011), 2882–
2886.
Li, J.X., Shi, F.B., Zhang, T., Wu, H.S., Sun, L.D. & Yan, C.H.: Ytterbium Stabilized Ordered Mesoporous
Titania for Near-Infrared Photocatalysis. Chem. Comm. 47, (2011), 8109–8111.
Mercado, C., Seeley, Z., Bandyopadhyay, A., Bose, S. & McHale, J.L.: Photoluminescence of Dense
Nanocrystalline Titanium Dioxide Thin Films: Effect of Doping and Thickness and Relation to Gas
Sensing. ACS Appl. Mater. Interfaces 2011, 3, 2281–2288.
Naya, S.I., Tanaka, M., Kimura, K. & Tada, H.: Visible-Light-Driven Copper Acetylacetonate Decomposition
by BiVO4 . Langmuir 27, (2011), 10334–10339.
Obregón, S., Caballero, A. & Colón, G.: Hydrothermal Synthesis of BiVO4 : Structural and Morphological
Influence on the Photocatalytic Activity. Appl. Catal. B: Environ. 117–118, (2012), 59–66.
Ohtani, B.: Preparing Articles on Photocatalysis: Beyond the Illusions, Misconceptions, and Speculation.
Chem. Lett. 37, (2008), 217–229.
Qin, W., Zhang, D., Zhao, D., Wang, L. & Zheng, K.: Near-Infrared Photocatalysis Based on
YF3 :Yb3+ ,Tm3+ /TiO2 Core/Shell Nanoparticles. Chem. Comm. 46, (2010), 2304–2306.
Reddy, B.M., Chowdhury, B. & Smirniotis, P.G.: An XPS Study of La2 O3 and In2 O3 Influence on the
Physicochemical Properties of MoO3 /TiO2 Catalysts. Appl. Catal. A 219, (2001), 53–60.
Silva, A.M.T., Silva, C.G., Dražic, G. & Faria, J.L.: Ce-doped TiO2 for Photocatalytic Degradation of
Chlorophenol. Catalysis Today 144, (2009), 13–18.
Song, X.C., Zheng, Y.F., Ma, R., Zhang, Y.Y., Yin, H.Y.: Photocatalytic Activities of Mo-doped Bi2WO6
Three-dimensional Hierarchical Microspheres. J. Hazard. Mater. 192, (2011), 186–191.
Spanier, J.E., Robinson, R.D., Zhang, F., Chan, S.W. & Herman, I.P.: Size-dependent Properties of CeO2−y
Nanoparticles as Studied by Raman Scattering. Phys. Rev. B, 64, (2001), 245407.
Tauc, J.: Absorption Edge and Internal Electric Fields in Amorphous Semiconductors. Mater. Res. Bull. 5,
(1970), 721–729.
Wang, X., Shan, G., Chao, K., Zhang, Y., Liu, R., Feng, L., Zheng, Q., Sun, Y., Liu, Y. & Kong, X.: Effects
of Er3+ Concentration on UV/blue Upconverted Luminescence and a Three-photon Process in the Cubic
Nanocrystalline Y2 O3 :Er3+ . Mater. Chem. Phys. 99, (2006), 370–374.
Wang, G., Qin, W., Zhang, J., Zhang, J., Wang,Y., Cao, C., Wang, L., Wei, G., Zhu, P. & Kim, R.: Enhancement
of Violet and Ultraviolet Upconversion Emissions in Yb3+ /Er3+ -codoped YF3 Nanocrystals. Optical
Mater. 31, (2008), 296–299.
Xu, A.W., Gao, Y. & Liu, H.Q.: The Preparation, Characterization, and their Photocatalytic Activities of
Rare-Earth-Doped TiO2 Nanoparticles. J. Catal. 207, (2002), 151–157.
Zhang, C. & Zhu, Y.: Synthesis of Square Bi2WO6 Nanoplates as High-Activity Visible-Light-Driven
Photocatalysts. Chem. Mater. 17, (2005), 3537–3545.
Zhang, L., Wang, W., Zhou, L. & Xu, H.: Bi2WO6 Nano- and Microstructures: Shape Control and Associated
Visible-Light-Driven Photocatalytic Activities. Small 3, (2007), 1618–1625.
Zhang, Z., Wang, W., Xu, J., Shang, M., Ren, J. & Sun, S.: Enhanced Photocatalytic Activity of Bi2WO6
Doped with Upconversion Luminescence Agent. Catal. Comm. 13, (2011), 31–34.
Zhou, T. Hu, J. & Li, J.: Er3+ Doped Bismuth Molybdate Nanosheets with Exposed {0 1 0} Facets and
Enhanced Photocatalytic Performance Appl. Catal. B: Environ. 110, (2011), 221–230
Zhu, K.R., Zhang, M.S., Chen, Q. & Yin, Z.: Size and Phonon-confinement Effects on Low-frequency
Raman Mode of Anatase TiO2 Nanocrystal. Phys. Lett. A 340. (2005), 220–227.
CHAPTER 18

Stabilized TiO2 nanoparticles on clay minerals for air


and water treatment

Elias Stathatos, Dimitrios Papoulis & Dionisios Panagiotaras

18.1 INTRODUCTION

TiO2 , among wide band gap semiconductors, is an efficient photocatalyst for organic compound
photodegradation both on aerated surfaces and in aqueous suspensions (Choi et al., 2007; Chong
et al., 2010). TiO2 is a relatively inexpensive semiconductor which exhibits high photocatalytic
activity, non-toxicity and stability in aqueous solutions, etc. (Choi et al., 2010; Sakkas et al.,
2010). Furthermore, the synthesis of mesoporous nanocrystalline anatase TiO2 particles, films
or membranes has extended their use in environmental remediation. Ultrafine TiO2 powders
with high particle surface area have good photocatalytic activity since reactions take place on the
surface of the photocatalyst. On the other hand, powders can easily agglomerate in larger particles
and as a consequence adverse phenomena to their photocatalytic activity are observed. Sol-gel
method, employing appropriate templating techniques or hydrothermal treatment of titanium
alkoxides are among common techniques used nowadays to fabricate efficient nanocomposite
TiO2 powders and films (Barbé et al., 1997; Stathatos et al., 2004). However, TiO2 is usually used
as disperse catalyst for its high catalytic surface area and activity (Litter, 1999). Nevertheless,
TiO2 powders cannot easily be recovered from aqueous systems when they are used for water
treatment. Highly dispersed TiO2 particles in suspension are difficult to handle and remove after
their application in water and wastewater treatment. Recently, many research studies have been
carried out to immobilize TiO2 catalyst onto various substrates as thin films and membranes (Li
et al., 2008; Wang et al., 2009). Despite of their lower catalytic surface area there is an increase
of the catalytic activity and utilization because of the extension of their field of applications.
The specific surface area, particle morphology and possible aggregation, phase composition and
number of -OH surface groups are among the most critical parameters for high photocatalytic
activity of the as-prepared immobilized powders and films. Another approach to enhance the
photocatalytic properties of the catalysts is the promotion of their porous structures.
Glass slides and fibers, membranes, activated carbon and zeolites are used as supports for TiO2
particles (An et al., 2008; Rose et al., 2009). The efficiency of photocatalytic procedure generally
decreases with catalyst immobilization as the catalyst mass and the illuminated total surface area
is lower than the case of pure TiO2 powder. However, the use of highly porous materials such as
clay minerals can be considered as alternative substrates for TiO2 immobilized particles. TiO2 in
pillared clays showed improved photocatalytic activities in decomposing air pollutants (e.g. NOx )
and organic substances in water (Chmielarz et al., 2009; Nikolopoulou et al., 2009; Paul et al.,
2012).
In aqueous dispersions, clay minerals have been used in combination with TiO2 to enhance the
removal of organic pollutants by photocatalytic processes. Furthermore, many clay minerals have
been used in combination with TiO2 in order to increase the photocatalytic activity in decomposing
NOx gas and volatile organic compounds such as toluene. Previous experiments showed that using
clay minerals with microfibrous or tubular morphology increased TiO2 photocatalytic activity.
Recent studies revealed that dispersing the TiO2 particles onto clay mineral’s surfaces under mild
conditions is a promising method to resolve the agglomeration problem of TiO2 (Papoulis et al.,
313
314 E. Stathatos, D. Papoulis & D. Panagiotaras

2013). Palygorskite clay mineral of hydrated magnesium aluminum silicate with lamellar structure
and halloysite with tubular nanostructure both possessing high surface areas could be considered
as suitable materials for TiO2 particles immobilization (Papoulis et al., 2013; Stathatos et al.,
2012). In this chapter we present the case of mesoporous nanocrystalline TiO2 in presence of
typical clay minerals such as palygorskite and halloysite, using (i) a template technique based on
the sol-gel method with surfactant molecules and (ii) typical hydrothermal treatment in relatively
low temperature. The enhanced photocatalytic activity of TiO2 nanoparticles in combination with
the advantages of palygorskite nanocomposite fibers as supports is presented to a typical example
for water treatment polluted with a textile. In particular, the discoloration of azo-dye Basic Blue 41
in aqueous solutions under the presence of composite palygorskite/TiO2 photocatalyst is presented
as a typical example for treatment of polluted water. The synergistic effect between TiO2 and clay
mineral fibers is also examined in spite of small amount of immobilized TiO2 catalyst. We have
chosen to present the case of an azo-dye as a target molecule because dye wastewater pollution
from textile industry is one of the major sources of environmental pollution introducing intense
coloring and toxicity to the aquatic system. Azo dyes are extensively used in textile industry and
they are very stable to ultraviolet and solar light irradiation. They are also resistant to biological
treatment and after reduction carcinogenic aromatic amines can be formed. Photocatalysis is
one of the methods that can be successfully applied to the oxidation and final removal of azo
dyes to the formation of carbon dioxide as a latter product. Moreover, we describe the synthesis,
characterization and photocatalytic activity of nanosized TiO2 particles supported on halloysite
nanotubular clay mineral as support by using hydrothermal treatment under mild conditions.
This method does not require stabilizing agents or calcination. We present an evaluation of their
photocatalytic efficiency against inorganic and organic air pollutants: (i) in decomposing NOx
gas, an important air pollutant which contributes to the formation of acid rain and to stratospheric
ozone destruction; (ii) in toluene vapor photo-elimination, one of the most prevalent and harmful
volatile organic compounds typical of urban atmospheres.

18.2 TIO2 NANOPARTICLES AND FILMS

18.2.1 Sol-gel method for nanoparticles and films


Sol-gel method has led to the synthesis of a great variety of materials, the range of which is
continuously expanding. Thus the simple incorporation of organic dopants as well as the formation
of organic/inorganic nanocomposites offers the possibility of efficient dispersion of functional
compounds in gels, it allows modification of the mechanical properties of the gels and provides
materials with very interesting optical properties. A typical sol-gel route for making oxide matrices
and thin films is followed by hydrolysis of alkoxides, for example, alkoxysilanes, alkoxytitanates,
etc. (Brinker and Scherer, 1990).
Hydrolysis ≡ Ti-OR + H2 O →≡ Ti-OH + ROH (18.1)
Polycondensation ≡ Ti-(OH) → -Ti-O-Ti- + H2 O (18.2)

where R is a short alkyl chain (e.g. ethyl, butyl, or isopropyl). Hydrolysis (18.1) produces highly
reactive hydroxide species Ti-OH, which, by inorganic polymerization, produce oxide, i.e. -Ti-
O-Ti-, which is the end product of the sol-gel process.
However, a review in literature reveals an increasing interest in another sol-gel route based on
organic acid solvolysis of alkoxides (Birnie and Bendzko, 1999; Wang et al., 2001). This second
method seems to offer substantial advantages in several cases and it is becoming the method of
choice in the synthesis of organic/inorganic nanocomposite gels. As it has been earlier found
by Pope and Mackenzie (Pope and Mackenzie, 1986) and later verified by others, organic (for
example, acetic or formic) acid solvolysis proceeds by a two-step mechanism which involves
intermediate ester formation (Ivanda et al., 1999). Simplified reaction schemes showing gel
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 315

formation either by hydrolysis or organic acid solvolysis are presented by the following reactions.
(Note that in these reactions only one metal-bound ligand is taken into account, while acetic acid
(AcOH) is chosen to represent organic acids in organic acid solvolysis):
≡ Ti-OR + AcOH →≡ Ti-OAc + ROH (18.3)
ROH + AcOH → ROAc + H2 O (18.4)
≡ Ti-OAc + ROH → ROAc+ ≡ Ti-OH (18.5)
≡ Ti-OR+ ≡ Ti-OAc → ROAc + -Ti-O-Ti- (18.6)

It is obvious that acetic acid solvolysis is more complicated than direct hydrolysis and polycon-
densation of previous reactions as several different possibilities may define different intermediate
routes to obtain the metal oxide. Reaction (18.3) is a prerequisite of the remaining three reac-
tions. Occurrence of reaction (18.4) would mean that water may be formed which may lead to
hydrolysis. Reaction (18.5) would create reactive Ti-OH which would form oxide, while reaction
(18.6) directly leads to oxide formation. The above possibilities have been demonstrated by var-
ious researchers by spectroscopic techniques. However, there still exists a lot of uncertainly and
there is no concrete model to describe a well-established procedure leading to oxide formation
by organic acid solvolysis. For this reason, more work needs to be carried out on these systems.
Aforementioned reactions reveal one certain fact. The quantity of acetic acid in solution will be
crucial in affecting intermediate routes. Thus, reaction (18.4) is possible only if an excess of
acetic acid is present. Also the quantity of acetic acid will define whether the solvolysis steps will
simultaneously affect all available alkoxide ligands or will leave some of them intact and subject
to hydrolysis reactions.
Although the formation of TiO2 is achieved either in the case of common sol-gel method with
hydrolysis and polycondensation of titanium alkoxides or in the case of organic acid solvoly-
sis the nanostructure of the oxide is not guaranteed. Agglomeration in larger particles causes
adverse phenomena to the photocatalytic activity of TiO2 . In usual sol-gel methods modified with
amphiphilic organic molecules, synthesis of TiO2 follows two possible strategies: (i) formation
of inorganic network onto self-organized organic molecules and (ii) formation of organic layer
around inorganic network in a sol, depending on mixing order, organic concentration and prop-
erties, and additives. Thus appropriate templating techniques could be applied for well dispersed
particles with reproduced particle size and shape. Usually the presence of non-ionic surfactants
such as Triton X-100, Tween series, etc. which bear different length chain of ether groups and
Pluronic copolymers are among some choices which play a crucial role in organizing the struc-
ture of the material and in creating well defined and reproducible nanophases (Choi et al., 2006).
This group of surfactants participates in the formation and organization of clusters in solution,
subsequently, organizing the structure of the ensuing gel. As a result transparent nanocrystalline
TiO2 has been presently deposited as thin film on different substrates among them clay minerals
is the topic that this chapter deals with.
Finally, high-temperature annealing, usually at 450–500◦ C, is necessary to remove organic
material needed to suppress agglomeration of TiO2 particles and reduce stress during calcina-
tion for making crack-free films with good adhesion on substrates. Besides, high-temperature
treatment of films promotes crystallinity of TiO2 particles and their chemical interconnection for
better photocatalytic activity. Low sintering temperature yields TiO2 nanocrystalline films with
high active surface area but relatively small nanocrystals with many defects and poor intercon-
nection, thus lower efficiency. High sintering temperature for TiO2 films is then the most efficient
method for the preparation of high performance photocatalysts.

18.2.2 Hydrothermal route for TiO2 nanoparticles and films


Hydrothermal method can be defined as a synthesis process of crystals which depends on the
solubility of minerals in hot water under high pressure. The crystal growth is performed in an
316 E. Stathatos, D. Papoulis & D. Panagiotaras

Figure 18.1. Autoclave vessel for composite photocatalyst preparation.

apparatus consisting of a steel vessel under pressure called autoclave (Fig. 18.1), in which a nutri-
ent is supplied along with water. Hydrothermal method is a facile route to prepare controllable
and well defined crystalline TiO2 particles in only one step (Rajamathi and Seshadri, 2002). Usu-
ally TiO2 nanoparticles are fabricated by the aqueous hydrolysis of a titanium alkoxide precursor.
Hydrothermal synthesis is normally conducted in the sealed autoclave under controlled tempera-
ture and/or pressure with the reaction in aqueous solution. The temperature can be elevated above
the boiling point of water, and the pressure can be determined by the temperature and amount of
the solution added to the autoclave chamber. It is then followed by autoclaving at temperatures
up to 240◦ C to achieve the desired nanoparticle size and crystallinity (anatase) (Barbé et al.,
1997). Furthermore, appropriate substrates could be present during hydrothermal route into the
autoclave and finally uniform films could be formed.
As a consequence the hydrothermal method is a simple route to preparation of TiO2 nanos-
tructures with controllable phases, sizes, exposed facets, shapes and hierarchical structures in
one step.

18.3 STABILIZED TIO2 PARTICLES WITH SOL-GEL METHOD ON CLAY MINERALS.


PALYGORSKITE CLAY MINERAL AS SUPPORT FOR TIO2 PARTICLES

18.3.1 Materials and methods


For sol-gel synthesis of the TiO2 nanoparticles, we have chosen the method described in Equa-
tions (18.3)–(18.6) as it was found that it better works with clay minerals dispersions. Therefore,
commercially available titanium tetraisopropoxide (TTIP), acetic acid (AcOH) and isopropanol
(i-PrOH), were purchased from Sigma-Aldrich. Triton X-100 (X-100, polyethylene glycol tert-
octylphenyl ether) nonionic surfactant was chosen as a pore directing agent for TiO2 nanostructure
(Aldrich). Palygorskite-rich samples (PAL) from Western Macedonia Greece were size fraction-
ated by gravity sedimentation to obtain fillers of less than 2 micrometers. Basic Blue 41 (BB-41)
is chosen as target molecule for photocatalytic discoloration experiments.
For the composite TiO2 /PAL sol synthesis a suitable amount of X-100 was homogeneously
dissolved in isopropanol (i-PrOH). Before adding alkoxide precursor, AcOH was added to the
solution for the esterification reaction with i-PrOH. Then, TTIP was added under vigorous stir-
ring. The molar ratio of the materials was optimized at X-100:i-PrOH:AcOH:TTIP = 1:68:6:1.
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 317

Scheme 18.1. Procedure for TiO2 /Clay mineral composite nanostructure via sol-gel method.

Palygorskite powder was mixed with previous solutions in various quantities following PAL/TiO2
weight ratio 0:1, 1:2, 1:1, 3:2, 2:1. Finally the sols were transferred at a rotary evaporator in
order to remove the solvent. The viscous sols were heat-treated for 2 hours at 500o C (PLF 110/30,
Protherm programmable furnace) to remove all the organic content.
A Bruker D8 Advance diffractometer with CuKα (λ = 1.5406 Å) radiation and Bragg-Brentano
geometry was used for X-ray diffraction (XRD) studies of the PAL/TiO2 catalyst. The mercury
intrusion curves of all samples were measured with a Quantachrome PoreMaster 60 Mercury
Porosimeter and the surface area, porosity, and pore size distribution were derived by differen-
tiating them. For the visual morphology of PAL/TiO2 nanostructure, an environmental scanning
electron microscope (FESEM Zeiss SUPRA 35VP) was used and inspect sample homogene-
ity. Absorption measurements of BB-41 sols were carried out with a Hitachi U-2900 UV-Vis
spectrophotometer. For the photocatalytic experiments a cylindrical reactor was used in all exper-
iments. Air was pumped through a gas inlet using a small pump to ensure continuous oxygen
supply to the reaction solution while it was simultaneously agitating. Illumination of samples was
performed using four black light fluorescent tubes of 4 W nominal power, which were placed
around the reactor. The whole construction was covered with a cylindrical aluminum reflector.
Cooling was achieved by air flow from below the reactor using a ventilator. The catalyst was
in the form of PAL/TiO2 powder in various w/w proportions. The total mass of catalyst was
15 mg/100 mL in order to avoid any screening effect. The intensity of radiation reaching the reac-
tor on the side facing lamps was measured with a Solar Light PMA-2100 UV-Photometer and
found equal to 0.9 mW cm−2 . The dye concentration was 2.5 × 10−5 M in the aqueous solution.
This dye is strongly adsorbed on TiO2 -palygorskite nanostructures. For this reason, we stored the
solution in the presence of the photocatalyst in the dark for an hour and all of the photocatalytic
results were obtained after equilibrium.

18.3.2 Photocatalyst characterization


As referred in the introduction, several procedures have been applied to the immobilization of
TiO2 with enhanced properties and photocatalytic activity. Glass slides and fibers, membranes,
activated carbon and zeolites are among supports used for TiO2 particles immobilization. How-
ever, the use of highly porous materials such as clay minerals can be considered as alternative
substrates for TiO2 immobilized particles. As an example we present palygorskite clay mineral of
hydrated magnesium aluminum silicate with lamellar structure and high surface area as it could
be considered as suitable material for TiO2 particles immobilization (Scheme 18.1). The chemical
analysis of the clay fraction of the palygorskite rich sample mostly includes SiO2 (53.17 wt%)
followed by MgO (12.76 wt%) and Fe2 O3 (7.73 wt%). Aluminum and titanium oxides can be also
found in lower contents. The composite PAL/TiO2 material at weight proportions as 0:1, 1:2, 1:1,
3:2, 2:1 were finally calcined at relatively high temperature (∼480◦ C) to remove any organic
substance.
The XRD patterns of all samples, presented in Figure 18.2, showed the strong reflection at
2θ = 8◦ referred to clay mineral and 2θ = 25.1◦ referred to anatase TiO2 . In particular, strong
reflection at 2θ = 8◦ is attributed to palygorskite proving the enhanced crystallinity of the clay
mineral. The peaks at 2θ = 13.7◦ , 16.3◦ , 19.8◦ , 20.7◦ represent the Si-O-Si crystalline layer in
318 E. Stathatos, D. Papoulis & D. Panagiotaras

Figure 18.2. XRD patterns for pure palygorskite, sol-gel prepared TiO2 and composite nanocrystalline
material.

the clay. The reflection (101) is referred to the TiO2 anatase form appear at 2θ = 25.1◦ . The two
basic reflections at 2θ = 8◦ and 2θ = 25.1◦ for palygorskite and TiO2 are maintained at all samples
with different intensity ratios because of the variable proportion between them. The grain size
for both palygorskite and TiO2 has been calculated from XRD patterns using Scherrer’s equation:
D = 0.9λ/(s cos θ), where λ is the wavelength of the X-ray and s is the full width (radians) at half
maximum (FWHM) of the signal. The crystallite size for TiO2 is calculated 14.5, 17.1, 15, and 16.7
nm for samples 0:1, 1:2, 1:1, 3:2, 2:1, respectively. All the peaks indicated that the crystal phase
of the materials containing TiO2 was anatase and the relative width of peaks indicated that the
size of the nanocrystallite was less than 17 nm. The palygorskite particle size was also calculated
to be from 23–35 nm for all samples. It should be noted that it is evident from the XRD patterns
that the calcination at 480◦ C for 15 min did not cause any phase transformation or distraction of
palygorskite. For all samples: pure palygorskite or mixtures of palygorskite and TiO2 particles,
the pore sizes are expected to have a broad range of length scales. Indeed, the total porosity
of the composite palygorskite-TiO2 material can be found ranging from 0.635–0.705 compared
to 0.540 of pure TiO2 nanoaparticles and 0.860 of pure palygorskite. Moreover, the specific
surface area can be found ranging from 66–100 m2 /g compared to 44.5 of pure TiO2 nanoparticles
and 123 of pure palygorskite. Finally the total pore volume of the composite material can be
0.77–1.21 cm3 g−1 compared to 0.45 of pure TiO2 nanoparticles and 3.14 of pure palygorskite.
From all presented data is obvious that palygorskite material is highly porous (85.8%). TiO2
particles are also porous (53.9%) while porosity of the composite material takes values between
those referred for palygorskite and TiO2 as it is expected. Moreover, high values for the specific
surface area of all samples can be measured. Among samples there is a specific weight ratio
between palygorskite and TiO2 (PAL/TiO2 = 3:2) where highest value for particle surface area
(100 m2 g−1 ) can be measured. Microscopically palygorskite microfibers after TiO2 treatment
according to the sol-gel method can be seen in Figure 18.3. Palygorskite samples usually contain
fibers in planar structures as it can be seen in Figure 18.3a. The average diameter of the fibers,
as they were observed before modification, is 40 nm while the length is between 500–2000 nm.
After modification, TiO2 nanoparticles uniform in size overlay palygorskite fibers. In Figure
18.3b palygorskite microfibers seem to be completely covered with uniform layers of TiO2 in
uniform particle distribution because of the organic template. The TiO2 crystal grains have a
spherical shape while they have an average size ranging from 13 to 16 nm after a closer view to
the inset image in Figure 18.3b. No cracks or peeling off traces around palygorskite boundaries
are seen. It is assumed that the organophilic interphase, assured by surfactant coating, acts like a
templating medium which provides titanium dioxide nanoparticles with relatively monodisperse
particle sizes on the surface of the clay mineral nanofibers.
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 319

Figure 18.3. FE-SEM images of pure palygorskite fibers (a) and modified with TiO2 particles via sol-gel
method (b). As an inset the composite material is presented in higher magnification.

18.3.3 Photocatalytic activity of sol-gel TiO2 modified palygorskite clay mineral for polluted
water with an azo dye
The photocatalytic activity of composite PAL/TiO2 could be evaluated in organic pollutants degra-
dation. As an example we present the photocatalytic activity of the nanocomposites to experiments
undertaken on Basic Blue 41 (BB-41) as a model azo dye pollutant in water.
Generally, it has been proposed that titanium dioxide mediated photodegradation involves the
generation of electron-hole pairs, which migrate to the photocatalyst surface forming surface
bound hydroxyl and superoxide radicals according to following equations:

TiO2 + hν → e− +
cb + hVB (18.7)
O2 + e−
CB → O−•
2 (18.8)
+
H2 O + hVB → HO• + H+ (18.9)

It is commonly known that the hydroxyl and superoxide radicals are the primary oxidizing
species in the photocatalytic process. These oxidative reactions result in the photodegradation of
the azo dye which initially starts with a discoloration of the dye and finally results to a complete
photodegradation. The various ratios of PAL/TiO2 show different photocatalytic activity as it is
expected. Photodiscoloration rate of BB-41 was calculated by the formula: A0 −A/A0 , where A0 is
the absorbance of the initial concentration of BB-41 solution and A is the final absorbance after
irradiation with UV light. Discoloration efficiency is determined as n (%) = A0 −A/A0 · 100%. For
the repeated use of the photocatalysts, the samples were washed with distilled water and dried
at 80◦ C while no further treatment was followed. The rate of discolorization was monitored with
respect to the change in intensity with time of the absorption peak at 610 nm. The absorption peak
of the dye diminished with time and disappeared during the reaction indicating that it had been
degraded. The results also show that there was no direct photolysis of BB-41 in the absence of
photocatalysts. In the case of the PAL/TiO2 weight ratio equal to 3:2 a complete discolorization
was reached within 90 min of illumination implied the synergistic effect between palygorskite
and TiO2 by preparing highly porous catalysts (Fig. 18.4). However, the further addition of
palygorskite (2:1) caused a slight decrease on the photocatalytic activity. We have also noticed
that there is a low photocatalytic response of palygorskite in UV light without the presence of
pure TiO2 which most probably due to the presence of light activated oxides that contains such
as Fe2 O3,TiO2 or the combination of them with MgO or CaO. The photocatalytic activity of all
samples appears in Figure 18.4.
As a consequence, the microfibrous nanocomposite materials can behave as photocata-
lysts acting separately from anatase nanoparticles, showing a synergistic effect. The enhanced
320 E. Stathatos, D. Papoulis & D. Panagiotaras

Figure 18.4. Photodiscoloration efficiency (%) of BB-41 with various PAL/TiO2 weight ratios.

performance of PAL/TiO2 = 3:2 ratio compared to the rest of the samples is attributed to the better
particle surface area. Degradation kinetics of BB-41 has been observed to follow first-order kinet-
ics and it is well established that photodiscoloration experiments follow Langmuir-Hinshelwood
model, where the global reaction rate r, is proportional to the surface coverage θ and CS total
number of sites (occupied or not) which is highly related to the total surface area of the adsorbent:
the greater the surface area the more sites and the faster the reaction, according to the following
equation:
dC k1 KC
r=− = k1 θCS = (18.10)
dt 1 + KC
where k1 is the reaction rate constant, K is the adsorption coefficient of the reactant, Cs represents
the maximum total coverage and C is the reactant concentration. Besides, it should be mentioned
that k1 >> k ads , kdes (adsorption, desorption rate constants respectively). In the case that C is very
small, KC factor is negligible in respect to unity and the equation describes first-order kinetics.
The integration of Equation (18.10) yields to the Equation (18.11):
 
C
− ln = kapp t (18.11)
C0

With limit condition that on t = 0 we have the initial concentration C0 · kapp is the apparent
first-order rate constant. The maximum value for rate constant (38 × 10−3 min−1 ) was calculated
for sample PAL/TiO2 = 3:2 while the value for pure TiO2 film was estimated at 18 × 10−3 min−1 .
Furthermore, all the samples exhibited better performance than pure TiO2 . This is attributed to
better structural characteristics of the palygorskite/TiO2 samples compared to pure TiO2 tabulated
at porosity and particle surface area. Finally, it has been found that the same photocatalyst can
be used in several photocatalytic cycles without any important loss to its efficiency. In particular,
three successive cycles of photocatalytic reactions employing the same catalyst (PAL/TiO2 = 3:2)
showed a total 8% efficiency loss.

18.4 STABILIZED TIO2 PARTICLES WITH HYDROTHERMAL ROUTE ON CLAY


MINERALS. HALLOYSITE CLAY MINERAL AS AN EXAMPLE

18.4.1 Materials and methods


In this section we present an example for clay mineral/TiO2 composite catalyst preparation using
hydrothermal route as an alternative method for this purpose. A stock TiO2 sol dispersion was
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 321

prepared mixing titanium tetraisopropoxide (TTIP), with hydrochloric acid, extra pure water
and absolute ethanol (Langlet et al., 2003). The TiO2 stock dispersion was diluted in absolute
ethanol to give a 0.05 M concentration of TTIP. Halloysite samples (HAL) were obtained from
Limnos Island, Greece, or Utah, USA. The samples were size fractionated to obtain sizes < 2 µm
by gravity sedimentation. Separation of the clay fraction was carried out using centrifugation
methods. The clay fractions of the most halloysite-rich samples were used for the preparation of
halloysite/TiO2 nanocomposites. The composite HAL/TiO2 photocatalyst was prepared according
to the following method: An aliquot of TiO2 sol was stirred for 2 h and the (1% w/w) halloysite-
water dispersion was added to the dispersion in order to obtain a final TiO2 content of 70% w/w.
The slurry was stirred for 24 h and the resulting dispersion was centrifuged for 10 min (3800 rpm)
followed by three times centrifuge washing with triple distilled water. The HAL/TiO2 composite
catalyst was then dispersed in a 1:1 water:ethanol solution, prior to hydrothermal treatment in
autoclave for 5 h at 180◦ C. The products were centrifuged for 15 min (at 3800 rpm) and dried
in an oven for 3 h (at 60◦ C). The phase compositions of untreated and TiO2 treated halloysite
samples were determined by X-ray diffraction (using a Bruker D8 advance diffractometer, with
Ni-filtered CuKα radiation). XRD patterns were obtained from oriented or random powder sam-
ples in a 2θ range of 2◦ to 60◦ at a scanning rate of 2◦ /min. HAL/TiO2 morphology and chemical
composition were examined with a SEM LEO SUPRA 35VP. The morphology of the samples was
also determined by transmission electron microscopy (TEM, JEOL 2010). Nitrogen adsorption-
desorption isotherms for each sample degassed at 100◦ C for 3 hours were obtained at 77 K using
Autosorb−1 (Quantachrome corporation). Brunauer-Emmet-Teller (BET) surface areas and pore
size distribution were determined from the isotherms. Pore size distribution of each sample was
obtained using density functional theory (DFT) method in which a N2 adsorption branch model
was selected.
The photocatalytic activity for nitrogen monoxide destruction was determined by measuring the
concentration of NO gas at the outlet of the reactor (373 cm3 of internal volume) during the light
irradiation of constant flowed 1 ppm NO-50 vol.% air mixed (balance N2 ) gas (200 cm3 min−1 ).
The photocatalyst (ca. 0.04 g) was placed in a hollow place of 20 mm × 15 mm × 0.5 mm on a
glass holder/plate and set in the center of the reactor. The photocatalyst under test was first equi-
librated with the flowing NO gas before turning on the light. A 450 W high-pressure mercury
arc lamp was used as the light source. The wavelength was controlled by selecting filters, i.e.,
Pyrex glass for > 290 nm, Kenko L41 Super Pro (W) filter > 400 nm and Fuji triacetyl cellu-
lose filter > 510 nm. The concentration of NO was determined using a NOx analyzer (Yanaco,
ECL-88A). For comparison, the photocatalytic reaction was also carried out using the standard
commercial TiO2 (Degussa P25). Activity and selectivity for the gas-phase photo-oxidation of
toluene were tested in a continuous flow annular photoreactor containing ca. 40 mg of pho-
tocatalyst as a thin layer coating on a Pyrex tube. The corresponding amount of catalyst was
suspended in 1 mL of ethanol, painted on a Pyrex tube (cut-off at ca. 290 nm) and dried at
room temperature. The reacting mixture (100 mL min−1 ) was prepared by injecting toluene
(Panreac, spectroscopic grade) into a wet (ca. 75% relative humidity) 20 vol.% O2 /N2 flow
before entering at room temperature to the photoreactor, yielding an organic inlet concentra-
tion of ca. 800 ppmv. After flowing the mixture for 3–4 h in the dark (control test), the catalyst
was irradiated by four lamps symmetrically positioned outside the photoreactor. The photo-
catalytic tests were performed under pure UV-light irradiation (Philips TL 6W/08) as well as
under a radiation spectrum simulating sunlight (Philips TL 6W/54-765). Reaction rates were
evaluated under steady state conditions, typically achieved after 4–5 h from the beginning of irra-
diation. No change in activity was detected for all samples within 24 h after reaching steady state
conditions. The concentration of reactants and products was analyzed using an on-line gas chro-
matograph (Agilent GC 6890) equipped with HP-PLOT-Q/HP-Innowax columns (0.5/0.32 mm
I.D. × 30 m) and TCD/FID detectors for the quantification of CO2 and organic substances
(toluene, photocatalytic products), respectively. The photocatalytic performance was conducted in
duplicate.
322 E. Stathatos, D. Papoulis & D. Panagiotaras

(a) (b)
Relative Intensity

(1)

(2)

Mag = 137.16 KX
10 20 30 40 50 60 100 nm
0

Figure 18.5. (a) XRD patterns of TiO2 /halloysite composites with two different samples of halloysite (1)
From USA (Utah) and (2) From Greece (Limnos Island). (b) SEM image of TiO2 /halloysite
composite (TEM image of the same sample as an inset), (Papoulis et al., 2013).

18.4.2 Photocatalyst characterization


Alternatively to sol-gel method, TiO2 stabilized particles on another clay mineral such as hal-
loysite could be applied via hydrothermal route. Figure 18.5a shows the XRD patterns of the
HAL/TiO2 nanocomposites which are characterized only by the presence of the minerals hal-
loysite and anatase confirming their purity. The basal reflections are significantly smaller than
anatase reflections while the XRD patterns of both nanocomposites exhibited all the characteristic
reflections of anatase at 25.3◦ , 37.9◦ , 47.6◦ and 54.8◦ 2θ degrees. It is therefore concluded that the
temperature (180◦ C) applied during the synthesis did not destroy the native dehydrated halloysite
mineral structure.
As in the case of PAL/TiO2 mixture the specific surface area of HAL/TiO2 composite material
exhibits high specific surface area (∼190 m2 g−1 ) which is very important to the fabrication of
efficient photocatalysts. SEM observations of nanocomposites revealed that many uniform TiO2
grains of about 10–30 nm were deposited on the halloysite tubes (Fig. 18.5b). The homogeneous
distribution of TiO2 grains on halloysite tubes is anticipated in order to improve the properties
and potential uses of the prepared nanocomposites. The deposited TiO2 nanoparticles were found
to be distributed very well on halloysite surfaces but not homogeneously.
TiO2 nanoparticles tend to cover the lumen of most halloysite tubes and appear to have been
deposited at least partially within the lumen as observed by TEM (Fig. 18.5b inset). TEM obser-
vations confirm that TiO2 particles are well dispersed on halloysite tubes but also revealed that
their grain size varied from 3–10 nm.

18.4.3 Photocatalytic activity of TiO2 modified halloysite clay mineral for air purification.
In combination to the use of PAL/TiO2 composite photocatalyst for water treatment the clay
mineral-TiO2 composite material could be proved that is also efficient to the air purification
from gases and volatile compounds. As an example we demonstrate the photocatalytic activity of
Halloysite/TiO2 composite catalyst in decomposing NOx gas and volatile toluene in gas phase.
The prepared HAL/TiO2 samples showed significantly higher activity for the NOx decomposition
under visible-light irradiation (λ > 510 nm), up to ∼9.4 times and under UV- Visible light irradi-
ation (λ > 290 nm), up to ∼1.7 times than that of the commercial TiO2 , Degussa-P25 (Fig. 18.6).
On a total TiO2 mass basis, HAL/TiO2 nanocomposites are exceptionally better than standard
i.e. TiO2 , up to 13.4 times under visible-light irradiation (λ > 510 nm), and up to 2.4 times under
UV-visible light irradiation (λ > 290 nm), than that of the commercial TiO2 , P25. Of course the
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 323

Figure 18.6. Photocatalytic activities in decomposing NOx gas by commercial P25, HAL/TiO2 under
different illumination sources.

results could be varied depending on the structural properties of clay mineral as substrate. How-
ever photocatalytic experiments performed with untreated halloysite samples as controls showed
no photocatalytic activity to the destruction of NOx .
In addition, steady state reaction rates of toluene gas-phase photocatalytic oxidation by
Halloysite/TiO2 catalyst obtained under pure UV and sunlight irradiation showed larger cat-
alytic activity compared to the commercial TiO2 , P25. In particular, reaction rates (mol s−1 g−1 )
of 10−8 to 2 × 10−8 can be observed for commercial TiO2 , P25, while comparable higher rates of
3 × 10−8 to 6 × 10−8 could be monitored for HAL/TiO2 composite catalyst. The reaction rates of
toluene photo-oxidation by the Halloysite/TiO2 sample were almost 3 times higher compared to
that of P25 under pure UV. Similar results could be obtained for sunlight exposure of the samples.
Catalytic tests carried out in the dark confirmed that the measured activities were fully attributable
to photo-induced processes.
The photocatalytic performance of the Halloysite/TiO2 nanocomposites was found to be even
better when compared to standard commercial TiO2 on the basis of TiO2 content. The reaction
products of the catalytic tests were also checked. Under the experimental conditions of the present
study, the reaction products detected during toluene photo-oxidation were CO2 and benzaldehyde,
in agreement with previous results using similar or other TiO2 -based systems. Both halloysite-
supported TiO2 samples showed higher formation of total oxidation products (62 to 66% CO2
yields) compared to the commercial TiO2 -P25 (29% CO2 yield). However, higher photocatalytic
activity could be obtained depending on higher specific surface area of the Halloysite samples.
Halloysite interparticle porosity has no positive effect on photocatalytic activity (even though
it results in an increase of specific surface area) while halloysite unblocked lumen indicates
poorer TiO2 dispersion on halloysite tubes, as corroborated by TEM micrographs. As in the
case of palygorskite, we found that halloysite in the composite photocatalyst gives a shift to the
absorption spectrum of TiO2 to longer wavelengths (data are not presented, Papoulis et al., 2013;
Stathatos et al., 2012). The phenomenon is more intense when the quantity of clay mineral is
increased because the pure clay mineral has a long tail in the visible. The extent of the absorption
of composite material in longer wavelengths may give some explanations about the activity of
the photocatalyst to the visible. However Halloysite could serve as a support which apparently
plays a complex role i.e., it may retard recombination of separated charges generated by UV
light irradiation in the crystal lattice of TiO2 , thereby increasing catalytic activity. Conversely,
halloysite may have a shielding effect, i.e. through scattering or absorbing a certain percentage
of the incoming UV light irradiation, decreasing thus the catalytic activity. It is evident that
dispersion of TiO2 on halloysite surfaces is highly effective for air decontamination (decomposing
NOx and toluene gas) through photocatalytic activity.
324 E. Stathatos, D. Papoulis & D. Panagiotaras

18.5 CONCLUSIONS

Highly porous nanostructured particles and films of composite clay mineral/TiO2 can be syn-
thesized via sol-gel method or hydrothermal route. Two common clay minerals with different
structural characteristics such as palygorskite and halloysite can be used for the above purpose.
Both composites exhibit enhanced structural properties including crystallinity, porosity and par-
ticle surface compared to pure clay minerals and TiO2 . The composite photocatalysts exhibit high
activity to the decomposition of aqueous or gas phase pollutants. In particular, we demonstrated
the photocatalytic activity of palygorskite/TiO2 composites to the discoloration of an azo dye in
water while halloysite/TiO2 composites were employed to the destruction of gas phase pollutants
such as NOx and volatile toluene in gas phase. Furthermore, the composites (Halloysite/TiO2 )
showed significantly higher photocatalytic activity in decomposing NOx gas under visible-light
(λ > 510 nm) compared to commercial TiO2 . This implies the role of the clay mineral as efficient
support or TiO2 stabilization while a synergistic effect could be taken place to the destruction of
contaminants.

ACKNOWLEDGEMENTS

The authors would like to acknowledge financial support from the European Union (Lead Market
European Research Area Network – LEAD ERA) and the Regional Authority of Western Greece
under the project “INDOOR ECOPAVING”. The project is implemented under the Operational
Programme “DEPIN 2007-2013” Priority Axis (PA) ’Digital convergence and entrepreneurship in
Western Greece’, action “Transnational Business Collaboration Western Greece” and is co-funded
by the European Union –European Regional Development Fund and National Resources (NSRF
2007–2013). Besides, the authors are thankful to Dr. Vassilios Dracopoulos, FORTH/ICE-HT,
for his help to the FE-SEM images and XRD diffractograms.

REFERENCES

An, T., Chen, J., Li, G., Ding, X., Sheng, G., Fu, J., Mai, B. & O’Shea, K.E.: Characterization and the
photocatalytic activity of TiO2 immobilized hydrophobic montmorillonite photocatalysts: Degradation of
decabromodiphenyl ether (BDE 209). Catal. Today 139 (2008), pp. 69–76.
Barbé, C.J., Arendse, F., Comte, P., Jirousek, M., Lenzmann, F., Shklover, V. & Grätzel, M.: Nanocrystalline
titanium oxide electrodes for photovoltaic applications. J. Am. Ceram. Society 80 (1997), pp. 3157–3171.
Birnie, D.P. & Bendzko, N.J.: H-1 and C-13 NMR observation of the reaction of acetic acid with titanium
isopropoxide. Mat. Chem. Phys. 59 (1999), pp. 26–35.
Brinker, C.J. & Scherer, G.W.: Sol-gel science, the physics and chemistry of sol-gel processing. Academic
Press, Inc, 1990.
Chmielarz, L., Piwowarska, Z., Kustrowski, P., Gil, B., Adamski, A., Dudek, B. & Michalik, M.: Porous
clay heterostructures (PCHs) intercalated with silica-titania pillars and modified with transition metals
as catalysts for the DeNOx process. Appl. Catal. B: Environ. 91 (2009), pp. 449–459.
Choi, H., Stathatos, E. & Dionysiou, D.D.: Synthesis of nanocrystalline photocatalytic TiO2 thin films
and particles using sol–gel method modified with nonionic surfactants. Thin Solid Films 510 (2006),
pp. 107–114.
Choi, H., Stathatos, E. & Dionysiou, D.D.: Photocatalytic TiO2 films and membranes for the development
of efficient wastewater treatment and reuse systems. Desalination 202 (2007), pp. 199–206.
Choi, H., Al-Abed, S.R., Dionysiou, D.D., Stathatos, E. & Lianos, P.: TiO2 -based advanced oxidation
nanotechnologies for water purification and reuse. In: Sustainability science and engineering, 2. Elsevier,
2010, pp. 229–254.
Chong, M.N., Jin, B., Chow, C.W.K. & Saint, C.: Recent developments in photocatalytic water treatment
technology: A review. Water Res. 44 (2010), pp. 2997–3027.
Ivanda, M., Music, S., Popovic, S. & Gotic, M.: XRD, Raman and FT-IR spectroscopic observations of
nanosized TiO2 synthesized by the sol–gel method based on an esterification reaction. J. Mol. Struct.
480–481 (1999), pp. 645–649.
Stabilized TiO2 nanoparticles on clay minerals for air and water treatment 325

Langlet, M., Jenouvrier, P., Kim, A., Manso, M. & Trejo-Valdez, M.: Functionality of aerosol-gel deposited
TiO2 thin films processed at low temperature. J. Sol-Gel Sci. Technol. 26 (2003), pp. 759–763.
Li, F., Sun, S., Jiang, Y., Xia, M., Sun, M. & Xue, B.: Photodegradation of an azo dye using immobilized
nanoparticles of TiO2 supported by natural porous mineral. J. Haz. Mater. 152 (2008), pp. 1037–1044.
Litter, M.I.: Heterogeneous photocatalysis: Transition metal ions in photocatalytic systems. Appl. Catal.
B:Environ. 23 (1999), pp. 89–114.
Nikolopoulou, A., Papoulis, D., Komarneni, S., Tsolis-Katagas, P., Panagiotaras, D., Kacandes, G.H., Zhang,
P., Yin, S. & Sato, T.: Solvothermal preparation of TiO2 /saponite nanocomposites and photocatalytic
activity. Appl. Clay Sci. 46 (2009), pp. 363–368.
Papoulis, D., Komarneni, S., Panagiotaras, D., Stathatos, E., Toli, D., Christoforidis, K.C., Fernández-García,
M., Li, H., Yin, S., Sato, T. & Katsuki, H.: Halloysite-TiO2 nanocomposites: Synthesis, characterization
and photocatalytic activity Appl. Catal. B: Environ. 132–133 (2013), pp. 416–422.
Paul, B., Martens, W.N. & Frost, L.R.: Immobilised anatase on clay mineral particles as a photocatalyst for
herbicides degradation. Appl. Clay Sci. 57 (2012), pp. 49–54.
Pope, E.J.A. & Mackenzie, J.D.: Sol-gel processing of silica: II. The role of the catalyst. J. Non-Cryst. Solids
87 (1986), pp. 185–198.
Rajamathi, M. & Seshadri, R.: Oxide and chalcogenide nanoparticles from hydrothermal/solvothermal
reactions. Curr. Opinion Solid State Mater. Sci. 6 (2002), pp. 337–345.
Rose, G., Echavia, M., Matzusawa, F. & Negishi N.: Photocatalytic degradation of organophosphate and
phosphonoglycine pesticides using TiO2 immobilized on silica gel. Chemosphere 76 (2009), pp. 595–600.
Sakkas, V.A., Islam, Md.A., Stalikas, C. & Albanis, T.A.: Photocatalytic degradation using design of
experiments: A review and example of the Congo red degradation. J. Hazard. Mater. 175 (2010),
pp. 33–44.
Stathatos, E., Lianos, P. & Tsakiroglou, C. Highly efficient nanocrystalline titania films made from
organic/inorganic nanocomposite gels. Micropor. Mesopor. Mater. 75 (2004), pp. 255–260.
Stathatos, E., Papoulis, D., Aggelopoulos, Ch. A., Panagiotaras, D. & Nikolopoulou, A.: TiO2 /palygorskite
composite nanocrystalline films prepared by surfactant templating route: Synergistic effect to the
photocatalytic degradation of an azo dye in water. J. Hazard. Mater. 211–212 (2012), pp. 68–76.
Wang, C., Deng, Z.X. & Li, Y.D.: The synthesis of nanocrystalline anatase and rutile titania in mixed organic
media. Inorg. Chem. 40 (2001), pp. 5210–5214.
Wang, X., Liu, Y., Hu, Z., Chen, Y., Liu, W. & Zhao, G.: Degradation of methyl orange by composite
photocatalysts nano-TiO2 immobilized on activated carbons of different porosities. J. Hazard. Mater. 169
(2009), pp. 1061–1067.
This page intentionally left blank
CHAPTER 19

Photodegradation of beta-blockers in water

Virender K. Sharma, Hyunook Kim & Radek Zboril

LIST OF ABBREVIATIONS

ATL Atenolol
PPL Propranolol
MET Metoprolol
WWTP Wastewater Treatment Plant
TOC Total Organic Carbon
DOC Dissolved Organic Carbon
EPR Electron Paramagnetic Resonance
SRFA Suwannee River Fulvic Acid
SRHA Suwannee River Humic Acid
NOFA Nordic Lake Fulvic Acid
NOHA Nordic Lake Humic Acid

19.1 INTRODUCTION

The continued exponential growth in human population has created a corresponding increase in
the demand for the earth’s limited supply of freshwater. Safe guarding our water resources is one
of the pressing issues of the present time (Chong et al., 2010; Köhler et al., 2012 ). In order to
resolve the global water shortage issue, recycling or reusing of treated water from a municipal
wastewater treatment plant (WWTP) has been promoted over the past decades (Grant et al., 2012).
However, the treated water recycling or reusing program has recently seen a setback as a result
of the occurrence of persistent organic contaminants. Therefore, the reuse or recycle programs of
treated municipal or industrial wastewater have been significantly affected (Arnold et al., 2012;
De la Cruz et al., 2012). Among these persistent organics, and emerging contaminants, residual
pharmaceuticals in the aquatic environment have been of a great concern over the past decade
because they can pose long-term risks, which include potential toxicity to aquatic organisms
and the disruption of endocrine systems of higher organisms (Andreozzi et al., 2004; De Witte,
et al. 2011). Pharmaceuticals are highly used by humans and also as veterinary medicines to
prevent diseases and to protect human or animal health. However, if unused pharmaceuticals
are improperly disposed, they can be released to the environment, and can pose a threat to the
environment. Even the pharmaceuticals administered to humans or animals can be released to
the environment, since they are only partially absorbed by a body and a significantly portion is
released to sewage and eventually to the environment.
Pharmaceuticals are generally found in ng L−1 in the aquatic environment due to dilution
effects, sorption to sediments, and biological and photochemical processes (Fatta-Kassinos et al.,
2011). However, local inputs such as discharge from a hospital may cause higher levels of phar-
maceuticals in water. Pharmaceuticals are biologically active molecules, which can interfere with
biological systems such as enzymes and receptors. Among different classes of pharmaceuticals,
β-blockers are among the most important class of pharmaceuticals. Concentrations of β-blockers
have been found up to µg L−1 in surface waters (Benotti et al., 2009; Martínez Bueno et al., 2012).
The levels of β-blockers may increase with the increase in the aging population of the world. High
327
328 V.K. Sharma, H. Kim & R. Zboril

concentrations of β-blockers might be ecotoxicologically harmful. For example, the values of


48-h LC50 of β-blockers to the aquatic organisms were calculated as sub ppm to tens ppm (Huggett
et al., 2002; Küster et al., 2010). Especially, a β-blocker, propranolol (1-(1-methylethylamino)-3-
(1-naphthyloxy)propan-2-ol, PPL) has been reported showing higher toxicity to aquatic organisms
compared to other β-blockers, atenolol (ATL) and metoprolol (MET) (Fent et al., 2006). Signif-
icantly, mixtures of β-blockers may also have additive effective to the organisms; contributing to
toxicity even at low concentrations.
Conventional treatment processes remove the beta-blockers incompletely in a WWTP, due to the
recalcitrant nature of β-blockers (Gabet-Giraud et al., 2010; Santiago-Morales et al., 2013). A few
studies have reported less than 20% removal of β-blockers including ATL, and PPL in a WWTP
(Gros et al., 2012). Biotransformation or biologically mediated removal of PPL has rarely been
reported. A very slow abiotic process removes only 20% of PPL, which has been known the only
lipophilic β-block with a bioaccumulation potential over 79 days (Barbieri et al., 2012; Maurer,
2007). Several advanced oxidation processes have been demonstrated to remove β-blockers from
water (Benitez et al., 2011; Chen et al., 2012; Isarain-Chávez et al., 2011; Köhler et al., 2012;
Sirés et al., 2010; Sirés and Brillas, 2012). Most of the studies have been carried out on the
treatment of PPL (Andreozzi et al., 2004; Benner and Ternes, 2009a; 2009b; Kim et al., 2009;
Piram et al., 2008; Romero et al., 2011; Sharma, 2008; Sirés and Brillas, 2012; Šojić et al., 2012).
Electrochemical approaches have also been suggested to remove β-blockers, e.g., PPL (Isarain-
Chávez et al., 2011; Sirés and Brillas, 2012). Importantly, the oxidative degradation of PPL
by environmentally friendly oxidant, ferrate(VI) (FeVI O2− 4 , Fe(VI)), has also been demonstrated
(Anquandah et al., 2013).
This chapter reviews the phototransformation of β-blockers in water. Initially, photodegradation
of β-blockers in water is presented, followed by the influence of water chemistry (pH, nitrate ion,
and organic matter) on the phototransformation rates. The mechanism of the reactions is also
briefly described.

19.2 PHOTOTRANSFORMATION IN WATER

UV spectra of β-blockers in water are shown in Figure 19.1 (Piram et al., 2008). Of the
several β-blockers, only few absorb light in the UV-C range (Fig. 19.1). Hence, photodegra-
dation of β-blockers by direct photolysis in the aquatic environment is limited only to small
numbers of β-blockers. Acebutolol, alprenolol, atenolol, pindolol, PPL and timolol can absorb
UV radiations over 250 nm to result in their phototransformation. Other β-blockers would require
the absorption of UV light of less than 250 nm to yield their degradation.
The direct photolysis has been studied under simulated sunlight using 1.1 kW xenon arc lamp
(Liu and Williams, 2007; Piramet et al., 2008). In this study, samples of varying concentrations
(0.3 µg L−1 to 10 mg L−1 ) were in deionized water. In the radiation system, the UV (<290 nm)
and the infrared irradiation (>800 nm) were eliminated using filters and the intensity was set to
the maximum level. The degradation kinetics of PPL, MET, and ATL in water followed first-
order kinetics (Liu and Williams, 2007; Piram et al., 2008). The first-order rate constant (kd(Xe) )
and calculated half-lives (t1/2(Xe) ) for the compounds are presented in Table 19.1. PPL could be
degraded via fast direct photolysis and the measured rate constants, kd(Xe) were between 0.033
and 0.058 h−1 (Table 19.1). The kinetic constant for the direct photolysis of MET was 5 orders
of magnitude smaller than that of PPL; the rate constants varied from 0.0007–0.0011 h−1 (Table
19.1). This suggests that chemical structures have influence on the kinetics of β-blockers. PPL
contains a naphthalene backbone while MET has a benzoic backbone. Comparatively, the values
of kd(Xe) for ATL were three times higher than those for MET (Table 19.1). The estimated values
of half-lives of the compounds in the Central Europe (C.EU) and the United States (U.S.) are also
given in Table 19.1. Half-life values indicate the role of weather differences between Europe and
the U.S. The light intensity is different for geographic locations and therefore the half-lives varied
in different regions.
Photodegradation of beta-blockers in water 329

Acebutolol Alprenolol Atenolol


)

)
⫺1

⫺1

⫺1
15000
.cm

.cm

.cm
5000 5000
10000
⫺1

⫺1

⫺1
ε (L.mol

ε (L.mol

ε (L.mol
5000

0 0 0
200 250 300 350 400 200 250 300 350 400 200 250 300 350 400
λ (nm) λ (nm) λ (nm)
5000
)

)
10000
⫺1

⫺1

⫺1
Metoprolol Nadolol Pindolol
.cm

.cm

.cm
⫺1

⫺1

⫺1
5000
2500 5000
ε (L.mol

ε (L.mol

ε (L.mol
0 0 0
200 250 300 350 400 200 250 300 350 400 200 250 300 350 400
λ (nm) λ (nm) λ (nm)
Propranolol 10000 Sotalol Timolol
)

)
⫺1

⫺1

⫺1
30000 5000
.cm

.cm

.cm
20000
⫺1

⫺1

⫺1
5000
ε (L.mol

ε (L.mol

ε (L.mol
10000

0 0 0
200 250 300 350 400 200 250 300 350 400 200 250 300 350 400
λ (nm) λ (nm) λ (nm)

Figure 19.1. β-blockers UV spectra for aqueous concentration of 10 mg L−1 (adapted from Piram, et al.,
2008 with the permission of Elsevier, Inc.).

Table 19.1. Measured and extrapolated direct photolysis kinetics of β-blockers in water.a (Adapted from
(Liu and Williams, 2007) with the permission of the American Chemical Society).

t1/2(solar) t1/2(solar) t1/2(solar) t1/2(solar)


Starting (d) June, (d) Dec., (d) June, (d) Dec.,
concentration kd(Xe) t1/2(Xe) C.EU C.EU U.S. U.S.
β-blocker (mg L−1 ) t (◦ C) (h−1 ) (h) 52◦ N 52◦ N 40◦ N 40◦ N

Propranolol 1 20 ± 0 0.033 21 1.0 12 1.9 6.8


25 ± 1 0.051 14 0.7 7.6 1.3 4.5
0.1 20 ± 0 0.036 19 1.0 11 1.8 6.3
22 ± 1 0.045 15 0.8 8.5 1.4 5.0
0.01 22 ± 1 0.045 15 0.8 8.5 1.4 5.0
0.001 23 ± 1 0.051 14 0.7 7.5 1.3 4.4
0.0003 23 ± 1 0.058 12 0.6 6.7 1.1 3.9
Average ± SD 16 ± 3 0.8 ± 0.2 8.7 ± 1.8 1.5 ± 0.3 5.1 ± 1.0
<12.5b 1.7 6.0
Metoprolol 10 24 ± 1 0.0007 990 44 449 95 298
1 24 ± 1 0.0011 630 28 286 61 190
Atenolol 0.1 26 ± 2 0.0019 365 16 166 35 110
0.01 24 ± 1 0.0020 347 15 157 33 104
200c 540
a Results are the means of two measurements and four analyses. Half-lives under solar irradiation were
extrapolated from the xenon results by assuming the quantum yields were independent of the wavelength
(Liu et al. 2004; Robinson et al. 2007). Light intensity data for Central EU and the U.S. were from Frank
et al. (1988) and Li et al. (1988), respectively. The first order direct photolysis rate constant (kd(Xe) ) under
the xenon lamp (295–800 nm). t1/2(Xe) is the half-life under the xenon lamp; t1/2(solar) is the half-life for near-
surface experimental conditions. b Results were measured values from (Andreozzi et al., 2003). Samples
were exposed to the sun in spring or summer in Italy. c Results were calculated from data given in (Andrisano
et al., 1999). Samples were exposed to a filtered 1.1 kW xenon arc lamp A filter system removed the UV
(<290 nm) and the infrared irradiation (>800 nm).
330 V.K. Sharma, H. Kim & R. Zboril

0.8 0.8
0.7 0.7
0.6 0.6

kobs (min⫺1)
0.5
0.5

Fraction
0.4
0.4
0.3
0.3
0.2
0.2 0.1
0.1 0.0
0.0 ⫺0.1
8 7 8 9 10
pH

Figure 19.2. pH Effects on photodegradation of atenolol (•) and metoprolol () (5.0 µM) in the presence
of FA (16 mg L−1 ), and the corresponding fraction of deprotonated forms of atenolol (•) and
metoprolol () (adapted from (Chen et al., 2012) with the permission of Elsevier, Inc.).

19.3 INFLUENCE OF WATER CHEMISTRY

19.3.1 pH
The effect of the solution pH on the photodegradation of ATL and MET in the presence of
fulvic acid (FA) and phosphate buffer is presented in Figure 19.2 (Chen, et al., 2012). The
photolysis experiments were performed using the 150-W xenon short arc lamp and the light of
wavelengths less than 300 nm was removed using filters to simulate sunlight. The pseudo-first-
order rate constants (kobs ) of the photodegradation did not show much difference over the pH in
the range from 6.0 to 8.0, but increased sharply with further increase in pH. This may be related
to the speciation of both ATL and MET, which have pK values of 9.6 and 9.7, respectively. The
fractions of deprotonated species of ATL and MET also increase in the pH range from 8.0 to 10.0
(Fig. 19.2), hence the deprotonated forms of the studied β-blockers may be more photoactive
than the protonated ones. Since the study was performed in the presence of FA, pH effect on the
speciation and the formation of FA in triplet state might also play a role in the observed trend;
the measured rate constants varied as a function of pH (Fig. 19.2).

19.3.2 Nitrate ion


The photolysis of ATL has been studied in the presence of nitrate ions. The results are shown in
Figure 19.3 (Ji et al., 2012). The kinetics of the photodegradation was first-order upon simulated
solar irradiation. As demonstrated in Figure 19.3, the photodegradation was dependent on the
concentration of nitrate and there was enhancement in the degradation rate as the concentration
of nitrate increased from 0.5 mmol L−1 to 10 mmol L−1 . The concentration of the reactive species,
HO• increases with increase in the concentration of nitrate ions (Brezonik and Fulkerson-Brekken,
1998) and hence the results shown in Figure 19.3 were expected. Pathways of photodegradation
of ATL are shown in Figure 19.4.
Three major pathways, which involved HO• have been suggested for the photodegradation
of ATL (Fig. 19.4). The electrophilic HO• attack on the aromatic ring resulted in the hydroxy-
lated ATL (IV) (Pathway I). A previous study reported that the hydroxylation was favored in the
cases of the aromatic pi electronic system and the electron-donating effect of ether side chain
(Song et al., 2008). In the second Pathway II, HO• first attack on the ipso-carbon to produce the
Photodegradation of beta-blockers in water 331

2.0

0.5 mmolL⫺1
1.6 1 mmolL⫺1
2 mmolL⫺1
5 mmolL⫺1
1.2 10 mmolL⫺1
ln(C0 /C)

0.8

0.4

0.0
0 60 120 180 240
Photolysis time (min)

Figure 19.3. Photolysis kinetics of ATL in the presence of different concentrations of nitrate ions : (×)
0.5 mM L−1 , (◦) 1 mM L−1 , () 2 mM L−1 , () 5 mM L−1 , (♦) 10 mM L−1 (adapted from Ji
et al., 2012 with the permission of Elsevier, Inc.).

hydroxycyclohexadienyl radicals or resonance-stabilized carbon-centered radicals. The cleav-


age of the ether side chain yielded the formation of 3-(isopropylamino)propane-1,2-diol (I) and
p-hydroxyphenylacetamide (Urano et al., 1996). The hydroxylated p-hydroxyphenylacetamide
product (II) was formed from the HO• electrophilic attack on p-hydroxyphenylacetamide. The
detachment of the amide moiety from the acetamide due to the attack of HO• and further oxidation
of the alkyl moiety by dissolved O2 formed the derivative of benzaldehyde. Finally, hydroxy-
lated 4-[2-hydroxy-3-(isopropylamino)propoxy] benzaldehyde (III) yielded by the adduct of HO•
(Fig. 19.4). Formation of hydroxylated products has also been observed in the UV irradiation of
PPL (Piram et al., 2012).

19.3.3 Types of natural organic matter


The presence of dissolved organic matter (DOM) in natural waters can also influence the pho-
todegradation of ATL (Wang et al., 2012; Zeng et al., 2012). In a study performed by Zeng
et al. (2012), photolysis using simulated sunlight source was applied for 10 µmol L−1 ATL in
the presence of 5–20 mg C L−1 Suwannee River Fulvic Acid (SRFA), Suwannee River Humic
Acid (SRHA), Nordic Lake Fulvic Acid (NOFA), and Nordic Lake Humic Acid (NOHA). The
results indicate that the DOM can mediate photodegradation of ATL under natural sunlit waters
(Fig. 19.5).
The influence of three inhibitors, i.e., bicarbonate ion, nitrate, and Fe(III) on the photodegra-
dation rate was also examined. The degradation rate decreased in the presence of bicarbonate,
therefore HO• was the main reactive species in the photodegradation of ATL (Fig. 19.5). Based
on the result, singlet oxygen and excited triplet of DOM as reactive species were ruled out. Nitrate
ions in solutions also could inhibit the photodegradation of ATL. This may be due to scavenging
of HO• by nitrate ions and photons absorption by nitrate ions. Iron species, which enhance pro-
duction of HO• through the photoprocess of Fe(III)–DOM complex, enhanced significantly the
photodegradation of ATL (Zeng et al., 2012).
332 V.K. Sharma, H. Kim & R. Zboril

OH
H
O N

H2N

Pathway III
Pathway I Pathway II HO•, -CHONH2
HO• OH HO• OH
H H
O N O N

H2N OH
OH H
O N O2
(IV)
O – OH OH
H
H2N O N

OH
H
O N
HO•
O
O H
OH
H2N H
O N

O
OH

(III)
OH
• • H
O • HO N
• •
O +

H2N (I)
H+, e–

OH
OH HO• O
O
H2N OH
H2N (II)

CO2, H2O, NH4+, NO3+

Figure 19.4. Possible photodegradation pathways of ATL in aqueous solution with nitrate (adapted from Ji
et al., 2013 with the permission of Elsevier Inc.).
Photodegradation of beta-blockers in water 333

0.0
SRFA SRHA

–0.3

–0.6

–0.9

–1.2
In(C/C0)

0.0
NOFA NOHA

–0.3

–0.6

–0.9

–1.2

0 10 20 30 40 50 0 10 20 30 40 50
t (hour)

Figure 19.5. Photodegradation of ATL in aqueous solution with DOMs of three different concentrations,
i.e., () 5 mg C L−1 , (•) 10 mg CL−1 , () 20 mg C L−1 ([ATL] = 10 µmol L−1 , pH = 6.8)
(adapted from Zeng, et al., 2012 with the permission of Elsevier Inc.).

19.4 MECHANISM

Numerous studies on the photodegradation of β-blockers in the presence of DOM (e.g. fulvic
acid) suggested that the phototransformation of β-blockers occur through the electron transfer
mechanism (Fig. 19.6). In a study performed with the simulated sunlight by (Chen et al., 2012),
it was found that interaction between β-blockers (e.g. ATL and MET) and triplet excited state
of fulvic acid (3 FA∗ ) would take place to yield the intermediate I and reduced species of FA
(process a). The intermediate II is formed through α-H electron transfer in the intermediate I
(process b). The formation of the carbon centered radicals was confirmed by the experiments
with electron paramagnetic resonance (EPR) (Liu and Williams, 2007). Process b subsequently
forms the intermediate III having C=N bonds (process c). The deisopropyl products are formed
from the decomposition of the intermediate III (process d). These photosensitized products of
ATL and MET have been confirmed by liquid chromatography-electron spray ionization-tandem
spectrometry (LC-ESI-MS/MS) (Chen et al., 2012).

19.5 MINERALIZATION AND TOXICITY

The toxicity of the phototransformation products has been studied using Daphnia magna, an
aquatic species (Ji et al., 2012). UV irradiation of ATL for 4 h resulted in ∼72% transformation
efficiency for ATL, while total organic carbon (TOC) removal efficiency was only ∼10%. The
result clearly demonstrated incomplete mineralization of ATL with the formation of intermediate
products. During the photolysis of ATL in water, decreased toxicity to D. magna was observed
after 24 h and 48 h exposure to the irradiated ATL. This suggests that the intermediates from the
334 V.K. Sharma, H. Kim & R. Zboril

a +• b
R O NH R O NH + FA (red)
FA/hν
OH OH
I

c d
R O NH R O NH R O NH2
+
OH OH OH
II III

* R = NH2COCH2- (atenolol); R = CH3OCH2- (metoprolol)

Figure 19.6. Photoproducts of ATL and MTL, and the proposed photosensitized degradation pathways in
FA solutions (adapted from Chen et al., 2012 with the permission of Elsevier, Inc.).

photolysis of ATL were less ecotoxic to freshwater species. The structural fragmentation of the
parent molecule may be the cause of the decreased toxicity (Andreozzi et al., 2004). Generally,
photodegradation was an effective process for decreasing the toxicity in the aquatic environment
(Ji et al., 2012).
In other study, PPL in deionized water (DIW) and river water (RW) were irradiated using a
solar simulator (λ = 295–800 nm) and comparative toxicity of PPL and its degraded mixtures were
tested using algae (Pseudokirchneriella subcapitata) and rotifers (Brachionus calyciflorus) (Liu
et al., 2009). Results showed decrease in the toxicity of photodegraded mixtures compared to that
of the parent molecules in all samples tested. Therefore, products from the phototransformation
of PPL were less toxic than PPL.

19.6 CONCLUSIONS

Studies on the photodegradation in aqueous solution using simulated sunlight in presence and
absence of photosensitizers indicates the transformation of β-blockers. The direct photolysis using
wavelength longer than UV-C is possible only for few β-blockers due to their characteristics for
UV spectra. Structures of β-blockers may control the phototransformations. The photodegradation
rate increases with increase in pH and the variation of species with pH influences the degradation
rate of each β-blocker. Nitrate ion enhances the photodegradation rate in the presence of fulvic
acid. The main reactive species, HO• , is largely responsible for the photodegradation of β-blockers
when DOM is also present in solution. The photodegradation products are formed through the
attack of the HO• to different moieties of β-blockers. Regarding the degradation mechanism,
initial step seems to be electron transfer, followed by several other steps to yield hydroxylated
products from the phototransformation of β-blockers in water. The products appear to be less
toxic to aquatic organisms than their parent molecule.

ACKNOWLEDGEMENTS

The authors thank the Center of Ferrate Excellence, Florida Institute of Technology for the support.
We would like to thank the grant from the R&D Program of MIKE/KEIT (10037331, Department
of Core Water Treatment Technologies based on the Intelligent BT-NT-IT Fusion Latform). We
thank anonymous reviewers for their comments which improved the chapter greatly.
Authors also thank Operational Program Research and Development for Innovations European
Regional Development Fund (project CZ 1.05/2.1.00/03.0058), and Technology Agency of the
Czech Republic “Competence Centers” (project TE01010218).
Photodegradation of beta-blockers in water 335

REFERENCES

Andreozzi, R, Marotta, R. & Paxéus, N.: Pharmaceuticals in STP effluents and their solar photodegradation
in aquatic environment. Chemosphere 50 (2003), pp. 1319–1330.
Andreozzi, R., Campanella, L., Fraysse, B., Garric, J., Gonnella, A., Lo Giudice, R., Marotta, R., Pinto, G. &
Pollio, A.: Effects of advanced oxidation processes (AOPs) on the toxicity of a mixture of pharmaceuticals.
Water Sci. Technol. 50 (2004), pp. 23–28.
Andrisano, V., Gotti, R., Leoni, A. & Cavrini, V.: Photodegradation studies on Atenolol by liquid
chromatography. J. Pharm. Biomed. Anal. 21 (1999), pp. 851–857.
Anquandah, G.A.K., Sharma, V.K., Panditi, V.R., Gardinali, P.R., Kim, H. & Oturan, M.A.: Ferrate(VI)
oxidation of propranolol: Kinetics and products. Chemosphere 91 (2013), pp. 105–109.
Arnold, R.G., Sáez, A.E., Snyder, S., Maeng, S.K., Lee, C., Woods, G.J., Li, X. & Choi, H.: Direct potable
reuse of reclaimed wastewater: It is time for a rational discussion. Rev. Environ. Health 27 (2012),
pp. 197–206.
Barbieri, M., Licha, T., Nödler, K., Carrera, J., Ayora, C. & Sanchez-Vila, X.: Fate of β-blockers
in aquifer material under nitrate reducing conditions: Batch experiments. Chemosphere 89 (2012),
pp. 1272–1277.
Benitez, F.J., Acero, J.L., Real, F.J., Roldan, G. & Casas, F.: Comparison of different chemical oxidation
treatments for the removal of selected pharmaceuticals in water matrices. Chem. Eng. J. 168 (2011),
pp. 1149–1156.
Benner, J. & Ternes, T.A.: Ozonation of metoprolol: Elucidation of oxidation pathways and major oxidation
products. Environ. Sci. Technol. 43 (2009a), pp. 5472–5480.
Benner, J. & Ternes, T.A.: Ozonation of propranolol: Formation of oxidation products. Environ. Sci Technol.
43 (2009b), pp. 5086–5093.
Benotti, M.J., Trenholm, R.A., Vanderford, B.J., Holady, J.C., Stanford, B.D. & Snyder, S.A.: Pharmaceu-
ticals and endocrine disrupting compounds in U.S. drinking water. Environ. Sci. Technol. 43 (2009),
pp. 597–603.
Brezonik, P.L. & Fulkerson-Brekken, J.: Nitrate-induced photolysis in natural waters: Controls on concne-
tration of hydroxyl radical photo-intermediates by natural scavenging agents. Environ. Sci. Technol. 32
(1998), pp. 3304–3310.
Chen, Y., Li, H., Wang, Z., Li, H., Tao, T. & Zuo, Y.: Photodegradation of selected β-blockers in aqueous
fulvic acid solutions: Kinetics, mechanism, and product analysis. Water Res. 46 (2012), pp. 2965–2972.
De la Cruz, N., Giménez, J., Esplugas, S., Grandjean, D., De Alencastro, L.F. & Pulgarín, C.: Degradation
of 32 emergent contaminants by UV and neutral photo-fenton in domestic wastewater effluent previously
treated by activated sludge. Water Res. 46 (2012), pp. 1947–1957.
De Witte, B., Van Langenhove, H., Demeestere, K. & Dewulf, J.: Advanced oxidation of pharmaceuticals:
Chemical analysis and biological assessment of degradation products. Crit. Rev. Environ. Sci. Technol.
41 (2011), pp. 215–242.
Fatta-Kassinos, D., Meric, S. & Nikolaou, A.: Pharmaceutical residues in environmental waters and
wastewater: Current state of knowledge and future research. Anal. Bioanal. Chem. 399 (2011),
pp. 251–275.
Fent, K., Weston, A.A. & Caminada, D.: Ecotoxicology of human pharmaceuticals. Aquat. Toxicol. 76 (2006),
pp. 122–159.
Frank, R. & Klopffer, W.: Spectral solar irrradiance in central Europe and adjacent North Sea. Chemosphere
17 (1988), pp. 985–994.
Gabet-Giraud, V., Miège, C., Choubert, J.M., Ruel, S.M. & Coquery, M.: Occurrence and removal of
estrogens and beta blockers by various processes in wastewater treatment plants. Sci. Total Environ.
408 (2010), pp. 4257–4269.
Gros, M., Rodríguez-Mozaz, S. & Barceló, D.: Fast and comprehensive multi-residue analysis of a broad
range of human and veterinary pharmaceuticals and some of their metabolites in surface and treated
waters by ultra-high-performance liquid chromatography coupled to quadrupole-linear ion trap tandem
mass spectrometry. J. Chromatogr. A 1248 (2012), pp. 104–121.
Huggett, D.B., Brooks, B.W., Peterson, B., Foran, C.M. & Schlenk, D.: Toxicity of select beta adrenergic
receptor-blocking pharmaceuticals (β-blockers) on aquatic organisms. Arch. Environ. Contam. Toxicol.
43 (2002), pp. 229–235.
Isarain-Chávez, E., Garrido, J.A., Rodríguez, R.M., Centellas, F., Arias, C., Cabot, P.L. & Brillas, E.:
Mineralization of metoprolol by electro-Fenton and photoelectro-Fenton processes. J. Phys. Chem. A 115
(2011), pp. 1234–1242.
336 V.K. Sharma, H. Kim & R. Zboril

Ji, Y., Zeng, C., Ferronato, C., Chovelon, J. & Yang, X.: Nitrate-induced photodegradation of atenolol in
aqueous solution: Kinetics, toxicity and degradation pathways. Chemosphere 88 (2012), pp. 644–649.
Kim, I.H., Yamashita, N., Kato, Y. & Tanaka, H.: Discussion on the application of UV/H2 O2 , O3 and O3 /UV
processes as technologies for sewage reuse considering the removal of pharmaceuticals and personal care
products. Water Sci. Technol. 59 (2009), pp. 945–955.
Köhler, C., Venditti, S., Igos, E., Klepiszewski, K., Benetto, E. & Cornelissen, A.: Elimination of pharmaceu-
tical residues in biologically pre-treated hospital wastewater using advanced UV irradiation technology:
A comparative assessment. J. Hazard. Mater. 239–240 (2012), pp. 113–120.
Küster, A., Alder, A.C., Escher, B.I., Duis, K., Fenner, K., Garric, J., Hutchinson, T.H., Lapen, D.R., Péry, A.,
Römbke, J., Snape, J., Ternes, T., Topp, E., Wehrhan, A. & Knackerk, T.: Environmental risk assessment of
human pharmaceuticals in the European union: A case study with the β-blocker atenolol. Integ. Environ.
Assess. Manage. 6 (2010), pp. 514–523.
Li, A.S.W., Cummings, K.B., RoethLing, H.P., Buettner, G.R. & Chignell, C.F.: A spin trapping database
implemented on the ABM/ PC/AT. J. Magn. Reson. 79 (1988), pp. 140–142.
Liu, Q. & Williams, H.E.: Kinetics and degradation products for direct photolysis of β-blockers in water.
Environ. Sci. Technol. 41 (2007), pp. 803–810.
Liu, Q.-T., Riddle, A.M., Robinson, P.F. & Gray, N.: Roles of partioning and phototransformation in predicting
the fate and movement of pharmaceuticals in UK and US rivers. Proceedings 4th International Conference
on Pharmaceuticals and EDCs in Water, Minneapolis, Minnesota, 13–15 October 2004, National Ground
Water Association, Westerville, OH, USA, 2004.
Liu, Q., Cumming, R.I. & Sharpe, A.D.: Photo-induced environmental depletion processes of β-blockers in
river waters. Photochem.Photobiol. Sci. 8 (2009), pp. 768–777.
Martínez Bueno, M.J., Ulaszewska, M.M., Gomez, M.J., Hernando, M.D. & Fernández-Alba, A.R.: Simul-
taneous measurement in mass and mass/mass mode for accurate qualitative and quantitative screening
analysis of pharmaceuticals in river water. J. Chromatogr. A 1256 (2012), pp. 80–88.
Maurer, H.H: Analytical toxicology. Anal.Bioanal.Chem. 388 (2007), p. 1311.
Piram, A., Salvador, A., Verne, C., Herbreteau, B. & Faure, R.: Photolysis of β-blockers in environmental
waters. Chemosphere 73 (2008), pp. 1265–1271.
Piram, A., Faure, R., Chermette, H., Bordes, C., Herbreteau, B. & Salvador, A.: Photochemical behaviour
of propranolol in environmental waters: The hydroxylated photoproducts. Int. J. Environ. Anal. Chem. 92
(2012), pp. 96–109.
Robinson, P.F., Liu, Q.-T., Riddle, A.M. & Murray-Smith, R.: Modelling the impact of direct phototrans-
formation on predicted environmental concnetrations (PEC) of propranolol hydrochlorite in UK and US
rivers. Chemosphere 66 (2007), pp. 757–766.
Romero, V., De La Cruz, N., Dantas, R.F., Marco, P., Giménez, J. & Esplugas, S.: Photocatalytic treatment
of metoprolol and propranolol. Catal. Today 161 (2011), pp. 115–120.
Santiago-Morales, J, Agüera, A., Gómez, M.D.M., Fernández-Alba, A.R., Giménez, J., Esplugas, S. & Rosal,
R.: Transformation products and reaction kinetics in simulated solar light photocatalytic degradation of
propranolol using Ce-doped TiO2 . Appl. Catal. B: Environ. 129 (2013), pp. 13–29.
Sharma, V.K.: Oxidative transformations of environmental pharmaceuticals by Cl2 , ClO2 , O3 , and Fe(VI):
Kinetics assessment. Chemosphere 73 (2008), pp. 1379–1386.
Sirés, I. & Brillas, E.: Remediation of water pollution caused by pharmaceutical residues based on
electrochemical separation and degradation technologies: A review. Environ. Int. 40 (2012), pp. 212–229.
Sirés, I., Oturan, N. & Oturan, M.A.: Electrochemical degradation of β-blockers. Studies on single and
multicomponent synthetic aqueous solutions. Water Res. 44 (2010), pp. 3109–3120.
Šojić, D., Despotović, V., Orčić, D., Szabó, E., Arany, E., Armaković, S., Illés, E., Gajda-Schrantz, K., Dombi,
A., Alapi, T., Sajben-Nagy, E., Palágyi, A., Vágvölgyi, C., Manczinger, L., Bjelica, L. & Abramović,
B.: Degradation of thiamethoxam and metoprolol by UV, O3 and UV/O3 hybrid processes: Kinetics,
degradation intermediates and toxicity. J. Hydrol. 472–473 (2012), pp. 314–327.
Song, W., Chen, W., Cooper, W.J., Greaves, J. & Miller, G.E.: Free-radical destruction of β-lactam antibiotics
in aqueous solution. J. Phys. Chem. A 112 (2008), pp. 7411–7417.
Urano, Y., Higuchi, T. & Hirobe, M.: Substrate-dependent changes of the oxidative O-dealkylation mech-
anism of several chemical and biological oxidizing systems. J. Chem. Soc. Perkin Trans. 2:6 (1996),
pp. 1169–1173.
Wang, L., Xu, H., Cooper, W.J. & Song, W.: Photochemical fate of beta-blockers in NOM enriched waters.
Sci. Total Environ. 426 (2012), pp. 289–295.
Zeng, C., Ji, Y., Zhou, L., Zhang, Y. & Yang, X.: The role of dissolved organic matters in the aquatic
photodegradation of atenolol. J. Hazard. Mater. 239–240 (2012), pp. 340–347.
CHAPTER 20

Final conclusions

Marta I. Litter, Roberto J. Candal & J. Martín Meichtry

In this book, examples of non-conventional, advanced oxidation technologies (AOTs) and


advanced reduction technologies (ARTs), as sustainable solutions for environmental treatments
are presented, selected from recognized groups of different countries. AOTs are well known for
their capacity for oxidizing and mineralizing almost any organic contaminant and transforming
several inorganic pollutants strongly resistant to conventional treatments, such as metals and met-
alloids. The AOTs presented in this book are mainly TiO2 photocatalysis (HP) and photo-Fenton,
but others like UV/H2 O2 , enzyme/H2 O2 , peracetic acid and UV photolysis are also exemplified.
Chapter 1 is a review showing the different approaches based on the use of sunlight as eco-
nomical and sustainable irradiation source. The chapter indicates that solar TiO2 photocatalysis
and photo-Fenton are low cost AOTs already developed at real scale and suitable to be combined
with biological processes. These technologies are especially convenient for the treatment of the
so-called emerging contaminants such as detergents, pharmaceutical products and its metabolites,
personal care products, flame retardants, antiseptics, fragrances, industrial additives, steroids and
hormones, among others. One solar AOT of particularly interest is solar photo-Fenton at neutral
pH using complexing agents, which uses minimal amounts of Fe and H2 O2 . An analysis of dif-
ferent solar photocatalytic reactors is performed, with the conclusion that Compound Parabolic
Concentrators (CPCs) are the more efficient systems. The different alternatives for coupling AOTs
and biological wastewater treatments are also presented.
Chapter 4 is another revision article presenting a comparative kinetic study between natu-
rally and artificially promoted photodegradation of phenolic and lactonic biocides, which can be
decomposed either by direct or by indirect photolysis. Direct photodecomposition processes can
be used whenever organic pesticides absorb sufficient light in the UV region (250–300 nm), but
several factors limit the efficiency of these photoprocesses in aquatic environments: the trans-
parency of natural waters, the pesticide molar extinction coefficient and solubility in water, the
solar light absorption and the water pH. In the case of the direct photolysis of aromatic pesticides,
the conversion yield has been found by the authors to be in the range from 18 to 99.5% for reaction
times of 40–70 min, according to the molecular structure and the nature of the solvent. On the
other hand, indirect photolysis uses the free solar light and a photosensitizer that, after excitation,
is able to react with the pesticide or generate reactive oxygen species (ROS) that decompose
the pollutant. From several possible sensitizers, rose bengal riboflavin was found to be the best
choice. On the other hand, Riboflavin, unavoidably present in practically all natural water courses
and lakes, gives a smaller degradation rate, but this problem can be compensated by prolonging
the irradiation time.
In line with the previous studies, Chapter 19 describes the photodegradation of pharmaceuticals
such as β-blockers using simulated sunlight in the presence and absence of photosensitizers. β-
blockers are biologically active molecules, which can interfere with biological systems. The
direct photolysis using wavelengths longer than those in the UVC range is possible only for few
β-blockers. On the other hand, when dissolved organic matter is present, hydroxyl radicals (HO• )
are largely responsible for the photodegradation of β-blockers.
ARTs are emerging technologies from which reductive TiO2 heterogeneous photocatalysis and
use of zerovalent iron nanoparticles are increasingly studied in last times. The use of zerovalent
iron particles has been successfully tested in batch and in-situ treatments, while TiO2 hetero-
geneous photocatalysis is especially investigated for the treatment of heavy metal ions. Three
337
338 M.I. Litter, R.J. Candal & J. Martín Meichtry

examples of these novel technologies are included in this book. One is related to the use of
Fe(0) nanoparticles for perchlorate removal and the other two are related to TiO2 /UVA catalyzed
reduction for elimination of arsenic or nitrate from water.
Use of iron nanosized particles is a very economic novel tool for removal of contaminants
from water. Chapter 7 shows the combined use of an ion exchange resin (A530E) and nZVI
for removal of perchlorate from water. Perchlorate was successfully reduced to chloride through
the application of Fe0 particles simultaneously with sodium chloride and ethanol regenerative
solution. Chemical regeneratives (NaCl or CH3 CH2 OH) were used to desorb perchlorate from
the resin and made possible the solubilization of the contaminant in the nZVI suspension. This
novel remediation system is rapid and economic, and uses small amounts of nZVI.
Chapter 2 describes the reduction of As(V) and As(III) species by TiO2 photocatalysis under
anoxic conditions as an innovative way for arsenic removal from water. Solid As(0) is produced,
with the subsequent immobilization of the pollutant. It is pointed out that the process at pH 3 is
very efficient for As (V or III) removal at low concentrations (e.g., 1 mg L−1 , a common value
found in natural polluted groundwaters) allowing to attain As levels lower than those established
by regulations in drinking water (<10 µg L−1 ).
Chapter 10 refers the abatement of nitrate in drinking water through a comparative study of
photocatalytic and conventional catalytic technologies. The efficiency of an economic catalyst
formed on a TiO2 support complemented with a mixture of Pd-In as metallic additives under
conventional conditions is compared with the efficiency under UV radiation. The purpose is
to avoid the formation of nitrite and NH+ 4 , undesirable reaction products. The conversion with
the photocatalytic system after 120 minutes of irradiation (around 40%) was equivalent to that
obtained in the dark, but the former presented better selectivity for N2 production, while the dark
system produced a large amount of nitrite.
Several chapters refer the preparation and NH4+ characterization of modified TiO2 materials
or other catalysts useful for AOTs/ARTs. The emphasis was put in increasing the photocatalytic
activity and in shifting the light absorption edge to higher wavelengths. For example, Chapter 17
shows the hydrothermal preparation of a nanosized (10 nm) highly active TiO2 doped with Er3+ ,
presenting improved photoactivity for phenol degradation. The enhancement was explained by
a dual-type mechanism, where Er3+ improves the electronic charge separation process, and an
additional small contribution of near infrared photons takes place due to an energy transfer process
from Er3+ to TiO2 .
Chapter 14 reports the use of laboratory prepared zeolites, NH4 -ZSM-11 and H-ZSM-11, as
supports for the in-situ generation of titania from titanium isopropoxide, followed by calcination.
The resulting catalytic systems were completely characterized and evaluated for their activity in the
photodecomposition of the insecticide dichlorvos, which resulted in degradation percentages close
to that of commercial P25. The advantage of these supported catalysts is the easy separation from
the aqueous system and the possibility of reuse. The addition of H2 O2 increased the photocatalytic
degradation of the insecticide.
Chapter 18 shows the synthesis of highly porous nanostructured particles and films of compos-
ite clay mineral/TiO2 via a sol-gel method or a hydrothermal route starting from two common clay
minerals, palygorskite and halloysite. All mixed clay/TiO2 samples exhibited enhanced structural
properties compared to pure clay minerals and TiO2 , and higher photocatalytic activity than TiO2 .
In particular, palygorskite/TiO2 composites showed high photocatalytic activity for the discol-
oration of an azo dye in water, while halloysite/TiO2 composites were employed in the destruction
of gas phase pollutants such as NOx (including photoactivity under visible irradiation) and
toluene.
Chapter 16 presents the synthesis of modified Ag/AgCl plasmonic photocatalysts by a facile
and effective method, exhibiting high photocatalytic activity in the visible for the treatment of
rhodamine B, 4-chlorophenol and/or Cr(VI). Several kinds of Ag/AgCl composite materials,
including Ag/AgCl core-shell sphere, Ag/AgCl@Cotton-fabric, Ag-AgCl/WO3 hollow sphere
and Ag-AgCl@TiO2 , and Ag-AgI/Fe3 O4 @SiO2 were prepared, characterized and tested. TEM
and SEM images confirmed the core-shell structure of these catalysts. The synthesized materials
Final conclusions 339

showed a high photocatalytic activity under visible light (λ > 420 nm) for the treatment of the
three indicated model pollutants.
Chapter 3 presents the preparation of materials based on tungstophosphoric acid (TPA) immo-
bilized over NH4Y zeolite by wet impregnation. The DRS results indicate that the band gap energy
values of the materials were similar to those reported for TiO2 . The photocatalytic efficiency of
the samples was evaluated using methyl orange; the highest photocatalytic activity corresponded
to the sample with the highest TPA to NH4Y ratio. According to the results, the materials exhibit
suitable textural and physicochemical properties to be used as catalysts in the photocatalytic
treatment of wastewater containing azo dyes.
Other chapters refer to Fenton or photo-Fenton like reactions using heterogeneous catalysts
containing iron or copper supported on different substrates. These different approaches were
explored in order to improve mineralization, increase the operation pH and facilitate catalyst
separation. For example, Chapter 5 describes the Fenton-like oxidation of organics using a home-
made Cu-chitosan/γ-Al2 O3 catalyst and H2 O2 . The complete oxidation of a phenol solution,
performed in a laboratory recirculating packed bed glass reactor, led to complete phenol and
H2 O2 consumption, together with 90% TOC conversion after four hours with a small excess of
H2 O2 without metal leaching.
Chapter 12 shows the efficiency of different iron-containing solids in heterogeneous photo-
Fenton processes: iron pillared clays (Fe-PILCs) prepared in the laboratory, commercial goethite
and nanoparticulate zerovalent iron (nZVI). Methylene blue, methyl orange and 2-chlorophenol
were used for degradation tests under photo-Fenton conditions, using a fluidized bed and a stirred
batch reactor. Different mechanisms involving the participation of Fe, either immobilized or in
solution, were proposed, depending on the structure of the solid and the model pollutants. In
comparison with goethite and nZVI, Fe-PILCs required lower H2 O2 consumption to attain high
2-CP removal, although with a higher level of Fe leaching.
Chapter 13 presents a comparative study about the use of Fe(III) and Cu(II) supported on
montmorillonite (MMT) as catalysts for the discoloration and mineralization of reactive Orange 16
by photo-Fenton process. Copper containing montmorillonite and bio-montmorillonite obtained
after removal of Cu(II) from water were also used to test their performance as catalyst in the same
reaction. Fe-MMT was very active for the discoloration of dyes in water by Fenton or photo-
Fenton like processes at pH 3.0 and 6.0. Cu-MMT was also active as catalyst at pH 3.0 and 6.0,
but only in photo-Fenton like processes, with a discoloration rate lower than that corresponding
to Fe-MMT. Leaching of Cu(II) is observed at pH 3.0 but not at pH 6.0.
The combination of UVC irradiation and H2 O2 is a well-known process for water purification
that does not need catalysts and may work at different pH values. Chapters 6 and 9 exemplify
the application of this process to the elimination of pesticides in water, trying to find the best
conditions to treat solutions containing typical concentrations of the herbicides glyphosate and
2,4-D and chlorpyrifos, respectively. Microtox assays indicated that the toxicity of the solutions
was significantly reduced after the treatment, demonstrating that it is not necessary to reach a
complete mineralization in order to obtain almost non-toxic end products.
Decomposition of H2 O2 to generate HO• can also be catalyzed by enzymes. Chapter 8
describes the discoloration of Direct Blue 273 using two technologies: enzymatic discoloration
and adsorption. In the first case, a crude extract of Soybean peroxidase was used, which catalyze
decomposition of H2 O2 to generate HO• . In the second case, adsorption on a synthetic resin, a
polyampholyte named Poly-(EGDE-MAA-2MI) was used. Both technologies resulted to be very
effective and economic.
Water disinfection is of great concern in the entire world, and large efforts are made in the
investigation of alternative disinfection/oxidation methods. In this book, Chapter 11 describes
the reduction of the viability of Bacillus subtilis spores over UV-irradiated TiO2 films in the gas
phase. Different TiO2 films were analyzed and compared with the reference material P25. A
methodology for kinetic modeling is proposed.
Chapter 15 is a brief account of the possibilities of peracetic acid and hydrogen peroxide as
additional optional choices. Five disinfection methods were compared: disinfection methods: (i)
340 M.I. Litter, R.J. Candal & J. Martín Meichtry

UV disinfection, (ii) H2 O2 disinfection, (iii) peracetic acid disinfection, (iv) peracetic acid + UV
disinfection and (v) H2 O2 + UV disinfection. The preliminary results suggest that peracetic acid
may be considered as an environmentally friendly good alternative for water treatment.
The works discussed along the different chapters of this book give an overview of the latest
developments in AOTs and ARTs and the trends followed by different research groups to improve
the efficiency of the processes. AOTs and ARTs are not universal solutions for water and gas
treatments, but they represent an alternative to completely eliminate or increase the biodegrad-
ability of several recalcitrant compounds. These technologies are especially attractive in countries
with high solar irradiation where photo-based technologies can be used, diminishing the energy
cost. Latin-American, African, South-European and several Asian developing countries, having
serious pollution problems, can take benefits from these technologies.
Sustainable Energy Developments

Series Editor: Jochen Bundschuh


ISSN: 2164-0645
Publisher: CRC Press/Balkema, Taylor & Francis Group

1. Global Cooling – Strategies for Climate Protection


Hans-Josef Fell
2012
ISBN: 978-0-415-62077-2 (Hbk)
ISBN: 978-0-415-62853-2 (Pb)

2. Renewable Energy Applications for Freshwater Production


Editors: Jochen Bundschuh & Jan Hoinkis
2012
ISBN: 978-0-415-62089-5 (Hbk)

3. Biomass as Energy Source: Resources, Systems and Applications


Editor: Erik Dahlquist
2013
ISBN: 978-0-415-62087-1 (Hbk)

4. Technologies for Converting Biomass to Useful Energy –


Combustion, gasification, pyrolysis, torrefaction and fermentation
Editor: Erik Dahlquist
2013
ISBN: 978-0-415-62088-8 (Hbk)

5. Green ICT & Energy – From smart to wise strategies


Editors: Jaco Appelman, Anwar Osseyran & Martijn Warnier
2013
ISBN: 978-0-415-62096-3

6. Sustainable Energy Policies for Europe – Towards 100% Renewable Energy


Rainer Hinrichs-Rahlwes
2013
ISBN: 978-0-415-62099-4 (Hbk)

7. Geothermal Systems and Energy Resources – Turkey and Greece


Editors: Alper Baba, Jochen Bundschuh & D. Chandrasekaram
2014
ISBN: 978-1-138-00109-1 (Hbk)

8. Sustainable Energy Solutions in Agriculture


Editors: Jochen Bundschuh & Guangnan Chen
2014
ISBN: 978-1-138-00118-3 (Hbk)
9. Advanced Oxidation Technologies – Sustainable Solutions for Environmental Treatments
Editors: Marta I. Litter, Roberto J. Candal & J. Martín Meichtry
2014
ISBN: 978-1-138-00127-5 (Hbk)

10. Computational Models for CO2 Geo-sequestration & Compressed Air Energy Storage
Editors: Rafid Al-Khoury & Jochen Bundschuh
2014
ISBN: 978-1-138-01520-3 (Hbk)
9 Series: Sustainable Energy Developments
9
Advanced Oxidation Technologies (AOTs) or Processes (AOPs) are relatively new and innovative

Litter, Candal & Meichtry


technologies to remove harmful and toxic pollutants. The most important processes among
them are those using light, such as UVC/H2O2, photo-Fenton and heterogeneous photocatalysis
with TiO2. These technologies are also relatively low-cost and therefore useful for countries
under development, where the economical resources are scarcer than in developed countries.

This book provides a state-of-the-art overview on environmental applications of Advanced


Oxidation Technologies (AOTs) as sustainable, low-cost and low-energy consuming treatments
for water, air, and soil. It includes information on innovative research and development on TiO2
photocatalytic redox processes, Fenton, Photo-Fenton processes, zerovalent iron technology,
and others, highlighting possible applications of AOTs in both developing and industrialized

Advanced Oxidation Technologies


countries around the world in the framework of “A crosscutting and comprehensive look at
environmental problems”.

The book is aimed at professionals and academics worldwide, working in the areas of water
resources, water supply, environmental protection, and will be a useful information source for
decision and policy makers and other stakeholders working on solutions for environmental
problems.

SUSTAINABLE ENERGY DEVELOPMENTS – VOLUME 9 ISSN 2164-0645

The book series addresses novel techniques and measures related to sustainable energy
developments with an interdisciplinary focus that cuts across all fields of science, engineering
and technology linking renewable energy and other sustainable materials with human society. It
addresses renewable energy sources and sustainable policy options, including energy efficiency
and energy conservation to provide long-term solutions for key-problems of industrialized, Advanced Oxidation Technologies
developing and transition countries by fostering clean and domestically available energy and,
concurrently, decreasing dependence on fossil fuel imports and reducing greenhouse gas Sustainable solutions for environmental treatments
emissions. Possible applications will be addressed not only from a technical point of view, but
also from economic, financial, social, political, legislative and regulatory viewpoints. The book
series aims to become a state-of-the-art source for a large group of readers comprising different
stakeholders and professionals, including government and non-governmental organizations and
institutions, international funding agencies, universities, public health and energy institutions,
and other relevant institutions.

SERIES EDITOR: Jochen Bundschuh

Editors: Marta I. Litter, Roberto J. Candal & J. Martín Meichtry

an informa business

You might also like