Anju Malik, Vinod Kumar Garg - Bioremediation For Sustainable Environmental Cleanup-CRC Press (2024)
Anju Malik, Vinod Kumar Garg - Bioremediation For Sustainable Environmental Cleanup-CRC Press (2024)
Environmental Cleanup
Editors
Anju Malik
Department of Energy and Environmental Sciences
Chaudhary Devilal University
Sirsa, Haryana, India
p,
p,
A SCIENCE PUBLISHERS BOOK
A SCIENCE PUBLISHERS BOOK
First edition published 2024
by CRC Press
2385 NW Executive Center Drive, Suite 320, Boca Raton FL 33431
and by CRC Press
4 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN
Reasonable efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers
have attempted to trace the copyright holders of all material reproduced in this publication and apologize to
copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been
acknowledged please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or
utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including
photocopying, microfilming, and recording, or in any information storage or retrieval system, without written
permission from the publishers.
For permission to photocopy or use material electronically from this work, access www.copyright.com or contact
the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. For works
that are not available on CCC please contact [email protected]
Trademark notice: Product or corporate names may be trademarks or registered trademarks and are used only for
identification and explanation without intent to infringe.
DOI: 10.1201/9781003277941
In the last century, rapid industrialization and urbanization have left a large quantity of a wide
range of pollutants emitted from both natural and anthropogenic sources. These pollutants have
detrimental effects on the flora, fauna and natural environment. The persistence and prevalence
of these pollutants virtually in all the ecosystems of Earth are of a major concern. Although a
large number of conventional physical and chemical methods are available for cleaning up the
environment, they are not sustainable and have inherent limitations. Hence, an alternate eco-
sustainable, bio-based and environment-friendly approach recognized as Bioremediation has
emerged. In bioremediation, biodiversity acts as a toolbox providing various processes/mechanisms
to eliminate, immobilize/stabilize, degrade or transform various hazardous contaminants into
innocuous and value-added products. This approach uses a vast array of biological agents, especially
bacteria (microbial remediation), fungi (mycoremediation), algae (phycoremediation), higher plants
(phytoremediation), biochar, nano-biomaterials, etc.
This book Bioremediation for Sustainable Environmental Cleanup has a compilation of
seventeen chapters contributed by eminent researchers from various countries located in diverse
geographical areas including Argentina, Canada, Germany, India, Pakistan, South Africa, the
United Kingdom and the United States of America. The book starts with a brief overview of
various bioremediation approaches used to clean up the polluted environment. It is followed by the
chapters on bioremediation strategies like Bioprecipitation, Bioaccumulation and Biosorption in
Chapters 2, 3 and 4. Further, the remediation of Polycyclic Aromatic Hydrocarbons using
phytoremediation and mycoremediation has been described in Chapters 5 and 6. The current
status and prospects of microbe-assisted bioremediation of pesticides have also been included in
Chapter 7. Chapter 8 focuses on the degradation of different pesticides by the bacterium Pseudomonas
putida. Chapters 9 to 13 extensively discuss the remediation of metal(loid)s. Chapter 14 presents
an overview of the potential of Water Hyacinth (Eichhornia crassipes) in remediating wastewater
discharged from the paper and pulp, textile and dairy industries. Furthermore, the applications
of novel materials viz. biohybrids, nano-biomaterials and graphitic Carbon Nitride (g-C3N4), in
environmental clean up of heavy metals, pesticides, dyes and other organic pollutants have been
addressed/covered in Chapters 15, 16 and 17.
Conclusively, this book comprehensively describes the state-of-the-art and potential of emerging
bioremediation approaches, which have been employed for sustainable environmental clean up of
various environmental pollutants such as metal(loid)s, polycyclic aromatic hydrocarbons, dyes,
pesticides, petroleum hydrocarbons, etc., by using bacteria, fungi, algae, higher plants and novel
materials like biohybrids, nano-biomaterials and graphitic Carbon Nitride (g-C3N4). The emphasis
throughout, however, is on sustainable environmental clean up. Each chapter of the book can stand
alone, contains updated information and is clearly illustrated with figures, pictures and tables in a
scientific way.
We firmly believe that this book will cater to the interest of researchers, academicians,
environmentalists, agriculturalists, extension workers, industrialists, students at undergraduate
and postgraduate levels, practising engineers, policymakers and other enthusiastic people who are
unreservedly devoted to sustainable environmental clean up of pollutants.
iv Bioremediation for Sustainable Environmental Cleanup
The editors have sincerely worked for about two years to carefully select, edit, revise and compile
the contributed manuscripts to the scope of the book. We have strived hard to ensure that this book
is free from any erroneous or deceptive information and any such mistake is completely inadvertent.
Further, we wholeheartedly thank all our contributors for readily accepting our invitation, timely
submitting their innovative, high quality and valuable chapters, and their cooperation at all stages
in making this voluminous and high-quality outcome a successful attempt. Last but not the least,
we also express our deep sense of gratitude to our family members for their unconditional support
during this long journey.
Anju Malik
Sirsa, India
Vinod Kumar Garg
Bathinda, India
Contents
Preface iii
1. Bioremediation: A Sustainable Approach for Environmental Cleanup 1
Bharti Singh, Anju Malik and Vinod Kumar Garg
2. Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 19
Samantha M. Wilcox, Catherine N. Mulligan and Carmen Mihaela Neculita
3. Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 40
Andrea Saralegui, M. Natalia Piol, Victoria Willson, Néstor Caracciolo, Silvia Ramos
and Susana Boeykens
4. Bioaccumulation and Biosorption: The Prospects and Future Applications 56
P.F. Steffi and P.F. Mishel
5. PAHs in Terrestrial Environment and their Phytoremediation 67
Sandip Singh Bhatti, Astha Bhatia, Gulshan Bhagat, Simran Singh,
Salwinder Singh Dhaliwal, Vivek Sharma, Vibha Verma, Rui Yin and Jaswinder Singh
6. Fungal Strategies for the Remediation of Polycyclic Aromatic Hydrocarbons 86
Nitu Gupta, Sandipan Banerjee, Apurba Koley, Aman Basu, Nayanmoni Gogoi,
Raza Rafiqul Hoque, Narayan Chandra Mandal and Srinivasan Balachandran
7. Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats: 109
Current Status and Prospects
Karen Reddy, Shisy Jose, Tufail Fayaz, Nirmal Renuka, Sachitra Kumar Ratha,
Sheena Kumari and Faizal Bux
8. Pseudomonas putida: An Environment Friendly Bacterium 125
Sneha S. Das and Gunderao H. Kathwate
9. General Aspects/Case Studies on Sources and Bioremediation 145
Mechanisms of Metal(loid)s
Manoj Kumar, Sushma K. Varma, Renju, Neeraj Kumar Singh and Rajesh Singh
10. Metal(loid)s Toxicity and Bacteria Mediated Bioremediation 167
Sushant Sunder, Anshul Gupta, Mehak Singla, Rohit Ruhal and Rashmi Kataria
11. Lead Induced Toxicity, Detoxification and Bioremediation 187
Shalini Dhiman, Arun Dev Singh, Isha Madaan, Raman Tikoria, Driti Kapoor,
Priyanka Sharma, Nitika Kapoor, Geetika Sirhindi, Puja Ohri and Renu Bhardwaj
12. Microalgal Bioremediation of Heavy Metals: An Integrated Low-cost 207
Sustainable Approach
Anubha Kaushik, Sharma Mona, Randhir Bharti and Sujata
vi Bioremediation for Sustainable Environmental Cleanup
1.1 Introduction
Both urbanization and industrialization have resulted in a rise of contaminated land and water. As
both these activities are growing at a rapid rate, they are creating stress on renewable as well as
non-renewable resources. As a result, environmental pollution has increased globally during the
last several decades. The expansion of the manufacturing and agricultural industries has led to an
increase in the release of a wide variety of xenobiotic substances into the environment. Excessive
release of hazardous waste has resulted in a lack of clean water and soil disturbances, restricting
agricultural output (Kamaludeen et al. 2003). Heavy metals, petroleum oil, pesticides, hydrocarbons
such as aliphatic, aromatic and polycyclic aromatic hydrocarbons, nitroaromatic compounds,
chlorinated hydrocarbons such as polychlorinated biphenyls (PCBs) and perchloroethylene,
organophosphorus compounds, organic solvents (phenolic compounds) and phthalates are
all possible contaminants in hazardous waste (Megharaj et al. 2011). The economically and
environmentally viable management of these wastes is a serious challenge around the world.
According to third-world network data, more than 450 million kg of hazardous waste is released
into the air, water and land on a global scale (Singh 2014). If these pollutants are discharged into
the environment without being correctly treated, they may cause serious health concerns as well as
the extinction of many species of fauna and flora. Various kinds of contaminants are carried into the
ground via air and water, resulting in a slew of major environmental and health issues around the
world (Boopathy 2000). Different types of technologies have been developed for waste disposal that
use high-temperature incineration, soil washing, adsorption, flocculation, landfilling, pyrolysis and
chemical decomposition etc. Despite their high potential for effectiveness, these techniques are also
inconvenient, expensive and undesirable.
Many researchers have increased their efforts to identify more environmentally friendly and
cost-effective alternatives for the use of hazardous chemicals and treatments to eliminate current
dangerous contaminants for sustainable cleanup of the environment. Several underdeveloped
countries cannot afford to build environmentally friendly and economically viable machinery for
1
Department of Energy and Environmental Sciences, Chaudhary Devilal University, Sirsa, Haryana, India.
2
Department of Environmental Science and Technology, School of Environment and Earth Sciences, Central University of
Punjab, Bhatinda, Punjab, India.
* Corresponding author: [email protected], [email protected]
2 Bioremediation for Sustainable Environmental Cleanup
waste handling. In such desperate situations, a few environmentally beneficial techniques can
overcome the restrictions associated with safe and cost-effective waste management technology.
Bioremediation approaches are increasingly being used as preferred methods for the detoxification
and biodegradation of pollutants in waste management and cleanup programs to revert contaminated
sites to their natural state, as they are cost-effective, natural environment friendly and long-term
solutions (Kumar et al. 2018).
Bioremediation is described as a process that uses living organisms, primarily microorganisms,
green plants and their enzymes, to remove, degrade, mineralize, transform and detoxify environmental
pollutants and hazardous components of waste into harmless or less toxic forms during the treatment
of contaminated sites in order to restore them to their original state (Azubuike et al. 2016). Pesticides,
polyaromatic hydrocarbons, halogenated petroleum hydrocarbons, nitroaromatic chemicals, metals
and industrial solvents have all been reduced in concentration and toxicity using the bioremediation
technique (Dua et al. 2002).
Microorganisms are used in bioremediation to immobilize or change the chemical structure of
pollutants found in soils, sediments, water and air, resulting in partial degradation, mineralization or
transformation of the molecule. However, most of the pollutants exhibit resistance to degradation,
which causes them to persist in the environment. Such accumulation may sometimes pose a serious
risk. Bioremediation techniques may be effectively set in polluted fields with the right use of
natural and engineered microorganisms. The treatment of polluted groundwater, soil, wetlands,
industrial wastes and sludge can all benefit from bioremediation approaches. Phytoremediation is
the process in which different types of plants are used in the bioremediation of contaminants. It
is a non-conventional, cost-effective and eco-friendly technology that utilizes plants to remove,
transform or stabilize a variety of contaminants located in water, sediments or soils (Prasad et al.
2001, Ladislas et al. 2012). This chapter aims to provide a brief overview of bioremediation and its
guiding principles, as well as the numerous bioremediation approaches, specifically the in-situ and
ex-situ remediation strategies, their benefits and drawbacks.
1.2 Bioremediation
Bioremediation is the “use of living organisms to clean up toxins from soil, water, or wastewater”
(EPA 2016). The utilization of living organisms to ameliorate the contaminated environment viz.
water, soil, etc., is referred to as bioremediation (Brar et al. 2006, Antizar-Ladislao 2010, Latha
and Reddy 2013). In other words, it is a technique for eliminating environmental contaminants
and restoring the natural ecosystem while also avoiding additional pollution. Various species
and their products, such as bacteria (bacterial bioremediation), fungi (mycoremediation), plants
(phytoremediation) and biomolecules generated from organisms, are engaged in bioremediation
processes (derivative bioremediation). Bioremediation can be aerobic (Wiegel and Wu 2000,
Bedard and May 1995) or anaerobic (Komancová et al. 2003). Bioremediation can occur naturally,
which is referred to as “natural attenuation” or it can take place artificially, which is referred to
as “biostimulation”. Note that not all toxins can be effectively removed with bioremediation.
Bioremediation aids humanity and society in dealing with harmful chemical pollutants that, if not
removed or rendered safe, may be damaging to human health and the environment.
as bioaugmentation involves the use of different cultured microorganisms to degrade soil and
water contaminants. Microorganisms must enzymatically attack contaminants and transform them
into harmless compounds for bioremediation to be efficient (Vidali 2001). Since environmental
conditions must be favorable for microbial development for bioremediation to be effective,
environmental parameters are regulated/monitored to accelerate microbial growth and deterioration.
Bioremediation procedures are often less expensive than traditional methods like incineration, and
certain contaminants may be treated on-site, lowering exposure hazards for cleanup workers and
possibly broader exposure due to transportation problems. Bioremediation is more widely accepted
than other approaches since it is based on natural attenuation. Most bioremediation systems are
operated under aerobic circumstances; however, operating one under anaerobic conditions may
allow microbial organisms to degrade compounds that are typically resistant to degradation (Colberg
and Young 1995).
In situ Ex situ
1.4.1.1 Bioaugmentation
Bioaugmentation is an effective method to increase the biodegradation of chemical compounds which
are added to the contaminated soil or water. When bioattenuation or biostimulation is ineffective
against resistant compounds, this strategy is thought to be effective. This approach is primarily used
to remediate aromatic and chlorinated hydrocarbon-contaminated municipal wastewater and soil.
Microorganisms from different cultures are brought together to enhance bioaugmentation efficiency.
The proper utilization of microbes which are naturally occurring can help to degrade and entirely
mineralize organic pollutants (Bender and Phillips 2004). Bioaugmentation-aided phytoextraction
with Plant Growth-Promoting Rhizobacteria (PGPR) or Arbuscular Mycorrhizal Fungi (AMF) is
also a potential option for metal-contaminated soil cleanup (Lebeau et al. 2008).
1.4.1.2 Biostimulation
The process of stimulating microbial enzymes for the bioremediation of diverse xenobiotic
chemicals is known as biostimulation, which involves adding soil nutrients, trace minerals, electron
acceptors or donors to a contaminated site until the optimum pH is reached (Li et al. 2010). If
natural degradation does not occur or occurs at a slow rate, the environment must be manipulated
in order to stimulate biodegradation and increase reaction rates. This method is primarily intended
to remediate fuel, hydrocarbon, metal contaminated sites, non-halogenated Volatile Organic
Compounds and Semi-Volatile Organic Compounds (VOCs, SVOCs) pesticides and herbicide
pollution of soil and groundwater. The availability of carbon, nutrients such as N and P, temperature,
available oxygen, soil pH and the kind and quantity of organic pollutant itself all influence the rate
of microbial turnover of chemical pollutants (Carberry and Wik 2001). There are several examples
of contaminants being biostimulated by indigenous microbes. Microorganisms have been observed
to convert trichloroethene and perchloroethylene to ethane in a short period when lactate is added
during biostimulation (Shan et al. 2010). In-situ biostimulation is predicted to become a dependable
and safe cleaning solution as scientific data accumulates via these types of field tests.
1.4.1.3 Bioventing
Bioventing is an in-situ approach in which different microorganisms are used to degrade organic
components adsorbed on soils in the unsaturated zone. This approach involves adding very little
oxygen to contaminated soil at low airflow rates, which is sufficient for successful microbial
biodegradation while minimizing pollutant volatilization and emission into the atmosphere (Atlas
and Philp 2005). In this technique, organic materials that are added to soil in the unsaturated zone
are biodegraded by the native microorganisms. TCE (trichloroethane), ethylene dibromide and
dichloroethylene are the most common volatile pollutants found at locations containing mid-weight
petroleum products (such as diesel and jet fuel) (Lee et al. 2006, Latha and Reddy 2013).
1.4.1.4 Bioslurping
Bioslurping is a technology that combines a number of procedures, including vacuum-enhanced
pumping, bioventing and soil vapor extraction, to remove contaminants from soil and groundwater
while also accelerating microbial biodegradation. One disadvantage of this strategy is the soil’s low
permeability, which reduces the pace at which oxygen is transferred and further suppresses microbial
activity. These methods are often used to remove volatile and semivolatile organic pollutants from
soil and liquids (Vidali 2001).
1.4.1.5 Biosparging
Biosparging also known as air sparging is an in-situ treatment method that involves injecting air
under pressure below the water table to clean polluted groundwater and raising subsurface oxygen
concentrations to enhance biodegradation in saturated and unsaturated soils. VOCs that have
polluted groundwater or soils in the saturated zone are treated using this method. Vapors that have
Bioremediation: A Sustainable Approach for Environmental Cleanup 5
been volatilized pass into the unsaturated zone, where they are vacuum-extracted by a soil vapor
extraction system (Hardisty and Ozdemiroglu 2005, Neilson and Allard 2008).
1.4.2.1 Biopiling
A biopile is one of the several bioremediation strategies for treating hydrocarbon-contaminated soil
that involves piling the dirt on top of an air distribution system and aerating it. Landfarming and
composting processes are combined in the biopiling approach and are mostly used in the treatment
of areas polluted with petroleum hydrocarbons. This whole setup includes a treatment bed and
nutrients, an aeration system, an underground irrigation system and a leachate collecting system
(Azubuike et al. 2016). For its cost-effectiveness and the ability to manage temperature, pH, and
nutrient conditions, this approach is increasingly being employed for bioremediation (Whelan et al.
2015).
1.4.2.2 Landfarming
Landfarming is a technique that includes dumping polluted material onto the soil surface and tilling
it to mix and aerate it (Harmsen et al. 2007, Maciel et al. 2009). Fuels, non-halogenated volatile
organic carbons, pesticides and herbicides are among the contaminants that this technology is meant
to handle. The idea is to encourage indigenous biodegradative bacteria and make it easier for them
to degrade pollutants aerobically. Landfarming appears to be limited to the treatment of 10–35 cm
of surface soil (Kumar et al. 2018). This technology is commonly used to clean up polluted areas
with aliphatic and polycyclic aromatic hydrocarbons, as well as PCBs (Silva-Castro et al. 2012).
1.4.2.3 Bioreactors
In this process, the pollutant is degraded in a reactor or container under controlled conditions. The
bioreactor technique is used to remediate soil or water that has been polluted by volatile organic
pollutants such as BTEX (benzene, toluene, ethylbenzene and xylene). A slurry bioreactor is a
vessel and apparatus which is used to increase the bioremediation rate of soil-bound and water-
soluble pollutants as a contaminated soil and biomass slurry (Kumar et al. 2011). Batch, continuous,
sequential batch biofilm, membrane, fluidized bed, biofilm and airlift bioreactors are among the
many types of bioreactors available globally. Before being placed in a bioreactor, the contaminated
soil must be pretreated or the contamination can be eliminated from the soil in different ways
(EPA 2000).
1.4.2.4 Biofilters
Biofilters are most commonly used to remove gaseous contaminants. The removal of gaseous
contaminants is accomplished using columns packed with microorganisms by converting unwanted
elements, such as CO2, H2O and cell mass, into harmless objects (Boopathy 2000). Biofilters use
a combination of basic techniques such as assimilation, adsorption, desorption and debasement of
gaseous phase contaminants. Microorganisms form a biofilm that adheres to the surface of a solid-
packed medium. The filter bed medium is often composed of natural materials with a wide surface
area in the zone and some supplement supplies. Biofilters are commonly used in water supply
systems to manage moisture and add nutrients (Gopinath et al. 2018).
6 Bioremediation for Sustainable Environmental Cleanup
aquatic species, Benzo(a)Pyrene (BaP), is known for its mutagenic, carcinogenic and teratogenic
characteristics (IARC 1983, Juhasz and Naidu 2000, Jennings 2012). In literature, ligninolytic and
non-ligninolytic strains of fungi with the ability to breakdown PAH have been documented. The
degradation of BaP by white-rot fungus has been the subject of recent research (Hadibarataa and
Kristanti 2012, Bhattacharya et al. 2014). White rot fungi such as Phanerochaete chrysosporium,
Trametes versicolor, Cirnipellis stipitaria and Pleurotus ostreatus can breakdown most PAHs
efficiently as a carbon source. The white rot fungus Phanerochaete chrysosporium has a remarkable
ability to degrade and/or mineralize high-molecular-weight PAHs, and its genome has around
150 Polymorphic Cytochrome P450 Enzymes (CYPs) (Yadav et al. 2006) and has the ability
to oxidize BaP to 3-hydroxybenzo[a]pyrene (Syed et al. 2010). These CYPs were inducible by
naphthalene, phenanthrene, pyrene and BaP. Aspergillus, the most prevalent species of soil-dwelling
fungi, may metabolize some PAHs and is frequently found in contaminated areas (Cerniglia and
Sutherland 2010).
now concentrated on the creation of a less expensive bioremediation strategy, however, physical
and chemical treatments are often costly. Bioremediation, which uses microorganisms capable of
digesting harmful substances, is regarded as a viable, ecologically beneficent and economically
advantageous technique for getting rid of PCBs (Dercova et al. 2015). PCB bioremediation relies
primarily on bacterial aerobic utilization of the pollutant molecules, while benzoic acids that can be
further degraded by specialized strains of bacteria or substances with lower toxicity are typically the
products of biphenyl dioxygenase-initiated degradation (Murínová et al. 2014). The microorganisms
used in the remediation/degradation of PCBs are given in Table 1.2.
Table 1.2. Polychlorinated biphenyls (PCBs) degrading microorganisms.
1.6 Phytoremediation
Phytoremediation is an effective biotechnology for dealing with metal and metalloid contamination
and has a distinct environmental application area (Gu 2018). Green technology for environmental
remediation plays a competitive role in transferring the soluble and bioavailable fractions
of hazardous metals and metalloids from the solution or adsorption phase into green plants for
accumulation, detoxification and stabilization (Yu and Gu 2007a,b, 2008a,b).
Phytoremediation is a non-conventional, cost-effective and eco-friendly technology that utilizes
plants to remove, transform or stabilize a variety of contaminants present in water, sediments
or soils (Prasad et al. 2001, Ladislas et al. 2012). Phytoremediation is the application of plants
for in-situ or ex-situ treatment/removal of heavy metals from contaminated soils, sediments and
water (Garbisu and Alkorta 2001, Shah and Daverey 2020) provides many benefits compared
to traditional approaches of contaminated land remediation such as lowered cost, maintenance,
exposure to workers, and is generally more aesthetically pleasing (Sharma and Yeh 2020). Plants
that have more metal-removal capacity through accumulation are known as hyperaccumulators.
These plants are used to remediate contaminants by the uptake or transpiration of contaminated
water (Cho-Ruk et al. 2006, Smolyakov 2012). Table 1.6 shows various types of aquatic plant
species capable of accumulating HMs. Plants absorb a significant number of toxic elements and
nutrients, but only a small portion of them are damaging, affecting plants at higher concentrations.
When the degree of pollution in plants rises, they are harmed or die. Various processes for treating
the water have been designed, for example, biological, physical and chemical, but they are costly
and only applicable to a small amount of wastewater (Rezania et al. 2015). As a result, an alternative
wastewater treatment procedure, phytoremediation has been proposed, in which diverse plants are
used to clean wastewater and eliminate hazardous contaminants.
Table 1.6. Various types of aquatic plant species capable of accumulating HMs.
1.6.1.2 Phytovolatilization
In this process, contaminants are uptaken by roots, translocated in upper plant parts and released
through the leaves in the volatile form into the atmosphere (USEPA 2000). This process involves the
12 Bioremediation for Sustainable Environmental Cleanup
diffusion of volatile pollutants through open stomata of the leaves in a less toxic form. This involves
the removal of the pollutants in a gaseous form and the particular pollutant removal in safer forms.
1.6.1.3 Phytostabilization
Phytostabilization is a process in which the stabilization or fixation of heavy metals occurs so that
proper absorption and precipitation take place mainly through the soil, sediment and sludge (USEPA
2000). In this process, contaminants are absorbed and collected by roots or precipitate inside the root
zone of plants (rhizosphere).
1.6.1.4 Phytofiltration/Rhizofiltration
Precipitation and absorption by plants from soil and water are the main mechanisms in rhizofiltration.
In this process, the contaminants are restricted only to the root system. Various heavy metals are
retained by the root system in rhizofiltration (USEPA 2000). In rhizofiltration, plant roots grow very
rapidly and require minimal time for decontamination (Sarkar et al. 2011).
1.6.1.5 Phytotransformation
Phytotransformation, also known as phytodegradation is the breakdown of organic pollutants
sequestered by plants through (1) metabolic processes inside the plant; or (2) the influence of
substances produced by the plant, such as enzymes (EPA 1998). It contributes to the removal of
organic pollutants such as chlorinated solvents, herbicides and the breakdown of complex organic
compounds into simple ones (EPA 2000). As plants do not contain active transporters, these organic
contaminants are absorbed through passive uptake. When the degradation of contaminants occurs in
the rhizosphere, the process is called rhizodegradation. Organics such as polyaromatic hydrocarbons
and polychlorinated biphenyls can be mineralized by rhizospheric bacteria. Furthermore, enzymatic
breakdown by enzymes released by specific plants and associated microbe species, such as
dehalogenase, nitro-reductase, laccase peroxidase and others, acts on dangerous xenobiotics
(Cherian and Oliveira 2005).
1.7 Conclusion
In the chapter, a brief account of the principle, types, applications, benefits and drawbacks of
bioremediation techniques have been discussed. As a result, one can see that compared to other
physical and chemical remediation approaches, it is an emerging, interdisciplinary, effective and
environmentally friendly remediation approach that is currently paving the way to a more promising
Bioremediation: A Sustainable Approach for Environmental Cleanup 13
future. Approaches to bioremediation largely depend on the concepts of in-situ and ex-situ technology.
In order to select the bioremediation method that would remove contaminants effectively, different
factors such as the presence of a particular microbial community, the bioavailability of pollutants
and environmental conditions have to be taken into consideration. As a result, improving our
understanding of microbial populations and how they interact with their specific environment and
pollutants, learning more about microbial genomics to increase their capacity for biodegradation
and evaluating the efficacy of new bioremediation strategies in the field will allow one to develop
more convenient bioremediation techniques.
References
Abbas, N., M. T. Butt, M. M. Ahmad, F. Deeba and N. Hussain. 2021. Phytoremediation potential of Typha latifolia
and water hyacinth for removal of heavy metals from industrial wastewater. Chem. Int. 7(2): 103–111.
Agrawal, N., V. Kumar and S. K. Shahi. 2021. Biodegradation and detoxification of phenanthrene in in-vitro and
in vivo conditions by a newly isolated ligninolytic fungus Coriolopsis byrsina strain APC5 and characterization
of their metabolites for environmental safety. Environ. Sci. Pollut. Res., 1–16.
Aggarwal, P. K., J. L. Means, R. E. Hinchee, G. L. Headington and A. R. Gavaskar. 1990. Methods to select chemicals
for in situ biodegradation of fuel hydrocarbons. Batt. Col. Div. Oh.
Ajaz, M., A. Elahi and A. Rehman. 2018. Degradation of azo dye by bacterium, Alishewanella sp. CBL-2 isolated
from industrial effluent and its potential use in decontamination of wastewater. J. of Water Reuse Desalin. 8(4):
507–515.
Akansha, K., A. N. Yadav, M. Kumar, D. Chakraborty and S. Ghosh Sachan. 2022. Decolorization and degradation of
reactive orange 16 by Bacillus stratosphericus SCA1007. Folia Microbiol. 67(1): 91–102.
Ali, H., E. Khan and M. A. Sajad. 2013. Phytoremediation of heavy metals—concepts and applications. Chemosphere.
91(7): 869–881.
Antizar-Ladislao, B. 2010. Bioremediation: working with bacteria. Elements. 6(6): 389–394.
Atlas, R. M. and J. Philp. 2005. Bioremediation. Applied Microbial Solutions for Real-world Environmental Cleanup.
ASM Press.
Azab, E. and A. K. Hegazy. 2020. Monitoring the efficiency of Rhazya stricta L. plants in phytoremediation of heavy
metal-contaminated soil. Plants. 9(9): 1057.
Azubuike, C. C., C. B. Chikere and G. C. Okpokwasili. 2016. Bioremediation techniques–classification based on site
of application: principles, advantages, limitations and prospects. World J. Microbiol. Biotechnol. 32(11): 1–18.
Bako, C. M., T. E. Mattes, R. F. Marek, K. C. Hornbuckle and J. L. Schnoor. 2021. Dataset describing biodegradation
of individual polychlorinated biphenyl congeners (PCBs) by Paraburkholderia xenovorans LB400 in presence
and absence of sediment slurry. Data in Brief. 35: 106821.
Barathi, S. and N. Vasudevan. 2001. Utilization of petroleum hydrocarbons by Pseudomonas fluorescens isolated
from a petroleum-contaminated soil. Environ. Int. 26(5-6): 413–416.
Barroso, G. M., J. B. dos Santos, I. T. de Oliveira, T. K. M. R. Nunes, E. A. Ferreira, I. M. Pereira, D. V. Silva and
M. de Freitas Souza. 2020. Tolerance of Bradyrhizobium sp. BR 3901 to herbicides and their ability to use
these pesticides as a nutritional source. Ecol. Indic. 119: 106783.
Bedard, D. L. and R. J. May. 1995. Characterization of the polychlorinated biphenyls in the sediments of Woods Pond:
evidence for microbial dechlorination of Aroclor 1260 in situ. Environ. Sci. Technol. 30(1): 237–245.
Bender, J. and P. Phillips. 2004. Microbial mats for multiple applications in aquaculture and bioremediation. Bioresour.
Technol. 94(3): 229–238.
Bhattacharya, S., A. Das, K. Prashanthi, M. Palaniswamy and J. Angayarkanni. 2014. Mycoremediation of Benzo
[a] pyrene by Pleurotus ostreatus in the presence of heavy metals and mediators. 3 Biotech. 4(2): 205–211.
Boopathy, R. 2000. Factors limiting bioremediation technologies. Bioresour. Technol. 74(1): 63–67.
Boyle, A. W., C. J. Silvin, J. P. Hassett, J. P. Nakas and S. W. Tanenbaum. 1992. Bacterial PCB
biodegradation. Biodegradation. 3(2): 285–298.
Brar, S. K., M. Verma, R. Y. Surampalli, K. Misra, R. D. Tyagi, N. Meunier and J. F. Blais. 2006. Bioremediation
of hazardous wastes—a review. Practice Periodical of Hazardous. J. Hazard. Toxic Radioact. Waste. 10(2):
59–72.
Briceño, G., K. Vergara, H. Schalchli, G. Palma, G. Tortella, M. S. Fuentes and M. C. Diez. 2018. Organophosphorus
pesticide mixture removal from environmental matrices by a soil Streptomyces mixed culture. Environ. Sci.
Pollut. Res. 25(22): 21296–21307.
Cao, H., C. Wang, H. Liu, W. Jia and H. Sun. 2020. Enzyme activities during Benzo [a] pyrene degradation by the
fungus Lasiodiplodia theobromae isolated from a polluted soil. Sci. Rep. 10(1): 1–11.
14 Bioremediation for Sustainable Environmental Cleanup
Carberry, J. B. and J. Wik. 2001. Comparison of ex situ and in situ bioremediation of unsaturated soils contaminated
by petroleum. J. Environ. Sci. Health, Part A. 36(8): 1491–1503.
Çelik, Ö. and E. Y. Akdaş. 2019. Tissue-specific transcriptional regulation of seven heavy metal stress-responsive
miRNAs and their putative targets in nickel indicator castor bean (R. communis L.) plants. Ecotoxicol. Environ.
Saf. 170: 682–690.
Cerniglia, C. E. 1984. Microbial metabolism of polycyclic aromatic hydrocarbons. Adv. Appl. Microbiol. 30: 31–71.
Cerniglia, C. E. and J. B. Sutherland. 2010. Degradation of polycyclic aromatic hydrocarbons by fungi. In Handbook
of Hydrocarbon and Lipid Microbiology.
Chang, F. C., C. H. Ko, M. J. Tsai, Y. N. Wang and C. Y. Chung. 2014. Phytoremediation of heavy metal contaminated
soil by Jatropha curcas. Ecotoxicol. 23: 1969–1978.
Chaussonnerie, S., P. L. Saaidi, E. Ugarte, A. Barbance, A. Fossey, V. Barbe, G. Gyapay, T. Bruls, M. Chevallier,
L. Couturat, S. Fouteau, D. Muselet, E. Pateau, G. N. Cohen, N. Fonknechten, J. Weissenbach and D. Le
Paslier. 2016. Microbial degradation of a recalcitrant pesticide: chlordecone. Front. Microbial. 7: 2025.
Cherian, S. and M. M. Oliveira. 2005. Transgenic plants in phytoremediation: recent advances and new
possibilities. Environ. Sci. Technol. 39(24): 9377–9390.
Chen, B. Y., C. M. Ma, K. Han, P. L. Yueh, L. J. Qin and C. C. Hsueh. 2016. Influence of textile dye and decolorized
metabolites on microbial fuel cell-assisted bioremediation. Bioresour. Technol. 200: 1033–1038.
Chen, S., Q. Hu, M. Hu, J. Luo, Q. Weng and K. Lai. 2011. Isolation and characterization of a fungus able to degrade
pyrethroids and 3-phenoxybenzaldehyde. Bioresour. Technol. 102(17): 8110–8116.
Cho-Ruk, K., J. Kurukote, P. Supprung and S. Vetayasuporn. 2006. Perennial plants in the phytoremediation of lead
contaminated soils. Biotechnol. 5: 1–4.
Colberg, P. J. S. and L. Y. Young. 1995. Anaerobic degradation of nonhalogenated homocyclic aromatic compounds
coupled with nitrate, iron, or sulfate reduction. Microbial Transformation and Degradation of Toxic Organic
Chemicals. 307330.
Cooney, J. J., S. A. Silver and E. A. Beck. 1985. Factors influencing hydrocarbon degradation in three freshwater
lakes. Microb. Ecol. 11(2): 127–137.
Daneshvar, N., M. Ayazloo, A. R. Khataee and M. Pourhassan. 2007. Biological decolorization of dye solution
containing Malachite Green by microalgae Cosmarium sp. Bioresour. Technol. 98(6): 1176–1182.
Dellagnezze, B. M., S. P. Vasconcellos, A. L. Angelim, V. M. M. Melo, S. Santisi, S. Cappello and V. M. Oliveira.
2016. Bioaugmentation strategy employing a microbial consortium immobilized in chitosan beads for oil
degradation in mesocosm scale. Mar. Pollut. Bull. 107(1): 107–117.
Dellamatrice, P. M., M. E. Silva-Stenico, L. A. B. D. Moraes, M. F. Fiore and R. T. R. Monteiro. 2017. Degradation
of textile dyes by cyanobacteria. Braz. J. Microbiol. 48: 25–31.
Dercova, K., K. Laszlova, H. Dudášová, S. Murinova, M. Balaščáková and J. Škarba. 2015. The hierarchy in selection
of bioremediation techniques: the potentials of utilizing bacterial degraders. Chemické Listy 109: 279–288.
Dua, M., A. Singh, N. Sethunathan and A. Johri. 2002. Biotechnology and bioremediation: successes and
limitations. Appl. Microbiol. Biotechnol. 59(2): 143–152.
de Lima Souza, H. M., L. D. Sette, A. J. Da Mota, J. F. do Nascimento Neto, A. Rodrigues, T. B. de Oliveira, L. A. de
Oliveira, H. D. Santos Barroso and S. P. Zanott. 2016. Filamentous fungi isolates of contaminated sediment in
the Amazon region with the potential for benzo (a) pyrene degradation. Water Air Soil Pollut. 227(12): 1–13.
Eid, E. M., T. M. Galal, N. A. Sewelam, N. I. Talha, S. M. and Abdallah. 2020. Phytoremediation of heavy metals by four
aquatic macrophytes and their potential use as contamination indicators: a comparative assessment. Environ
Sci. Pollut. Res. 27: 12138–12151.
EPA. 1998. A Citizen’s Guide to Phytoremediation. EPA 542-F-98-011. U.S. Environmental Protection Agency,
Washington.
EPA. 2000. A citizen guide to phytoremediation. EPA 542-F-98-011. United States Environmental Protection Agency,
p.6. Available http//www.bugsatwork.com/XYCLONYX/EPA_GUIDES/PHYTO.PDF.
EPA. 2016. United States Environmental Protection Agency. https://www3.epa.gov/ (Accessed May 2016).
Eslami, N., A. Takdastan and F. Atabi. 2022. Biological Remediation of Polychlorinated Biphenyl (PCB)-Contaminated
soil using the vermicomposting technology for the management of sewage sludge containing Eisenia fetida
earthworms. Soil Sediment Contamin. An International Journal, 1–17.
Esparza-Naranjo, S. B., G. F. da Silva, D. C. Duque-Castaño, W. L. Araújo, C. K. Peres, M. Boroski and R. C.
Bonugli-Santos. 2021. Potential for the biodegradation of atrazine using leaf litter fungi from a subtropical
protection area. Curr. Microbiol. 78(1): 358–368.
Gangola, S., G. Negi, A. Srivastava and A. Sharma. 2015. Enhanced biodegradation of endosulfan by Aspergillus
and Trichoderma spp. isolated from an agricultural field of tarai region of Uttarakhand. Pestic. Res. J. 27(2):
223–230.
Gangola, S., A. Sharma, P. Bhatt, P. Khati and P. Chaudhary. 2018. Presence of esterase and laccase in Bacillus subtilis
facilitates biodegradation and detoxification of cypermethrin. Sci. Rep. 8(1): 1–11.
Bioremediation: A Sustainable Approach for Environmental Cleanup 15
Garbisu, C. and I. Alkorta. 2001. Phytoextraction: a cost-effective plant-based technology for the removal of metals
from the environment. Bioresour. Technol. 77(3): 229–236.
Germain, J., M. Raveton, M. N. Binet and B. Mouhamadou. 2021. Potentiality of Native Ascomycete strains in
bioremediation of highly polychlorinated biphenyl contaminated soils. Microorganisms. 9(3): 612.
Gopinath, M., R. H. Pulla, K. S. Rajmohan, P. Vijay, C. Muthukumaran and B. Gurunathan. 2018. Bioremediation of
volatile organic compounds in biofilters. In Bioremediation: Applications for Environmental Protection and
Management. Springer, Singapore, 301–330.
Gorbunova, T. I., D. O. Egorova, M. G. Pervova, T. D. Kyrianova, V. A. Demakov, V. I. Saloutin and O. N. Chupakhin.
2021. Biodegradation of trichlorobiphenyls and their hydroxylated derivatives by Rhodococcus-strains. J.
Hazard. Mater. 409: 124471.
Govarthanan, M., S. Fuzisawa, T. Hosogai and Y. C. Chang. 2017. Biodegradation of aliphatic and aromatic
hydrocarbons using the filamentous fungus Penicillium sp. CHY-2 and characterization of its manganese
peroxidase activity. RSC Adv. 7(34): 20716–20723.
Gu, J. D. 2018. Bioremediation of toxic metals and metalloids for cleaning up from soils and sediments. Appl.
Environ. Biotechnol. 3(2): 48–51.
Hadibarata, T. and R. A. Kristanti. 2012. Fate and cometabolic degradation of benzo [a] pyrene by white-rot fungus
Armillaria sp. F022. Bioresour. Technol. 107: 314–318.
Hanafiah, M. M., M. F. Zainuddin, N. U. Mohd Nizam, A. A. Halim and A. Rasool 2020. Phytoremediation of
aluminum and iron from industrial wastewater using Ipomoea aquatica and Centella asiatica. Appl. Sci. 10(9):
3064.
Hardisty, P. E. and E. Ozdemiroglu. 2005. The economics of groundwater remediation and Protection, CRC Press,
Boca Raton, FL.
Haritash, A. K. and C. P. Kaushik. 2009. Biodegradation aspects of polycyclic aromatic hydrocarbons (PAHs): a
review. J. Hazard. Mater. 169(1-3): 1–15.
Harmsen, J., W. H. Rulkens, R. C. Sims, P. E. Rijtema and A. J. Zweers. 2007. Theory and application of landfarming
to remediate polycyclic aromatic hydrocarbons and mineral oil contaminated sediments; beneficial reuse. J.
Environ. Qual. 36: 1112–1122.
Hernández-Vega, J. C., B. Cady, G. Kayanja, A. Mauriello, N., Cervantes, A., Gillespie, L. Lavia, J. Trujillo, M. Alkio
and A. Colón-Carmona. 2017. Detoxification of polycyclic aromatic hydrocarbons (PAHs) in Arabidopsis
thaliana involves a putative flavonol synthase. J. Hazard. Mater. 321: 268–280.
Horváthová, H., K. Lászlová and K. Dercová. 2018. Bioremediation of PCB-contaminated shallow river sediments:
the efficacy of biodegradation using individual bacterial strains and their consortia. Chemosphere.
193: 270–277.
IARC (International Agency for Research on Cancer) 1983. Polynuclear aromatic compounds, part 1, chemical,
environmental, and experimental data. IARC Monographs on the Evaluation of the Carcinogenic Risk of
Chemicals to Man. IARC Sci. Pub. 32: 33–451.
Jaiswal, S., D. K., Singh and P. Shukla. 2022. Lindane bioremediation by Paenibacillus dendritiformis SJPS-4, its
metabolic pathway analysis and functional gene annotation. Environ. Technol. Innov. 27: 102433.
Jennings, A. A. 2012. Worldwide regulatory guidance values for surface soil exposure to carcinogenic or mutagenic
polycyclic aromatic hydrocarbons. J. Environ. Manage. 110: 82–102.
Jiang, J., H. Liu, Q. Li, N. Gao, Y. Yao and H. Xu. 2015. Combined remediation of Cd–phenanthrene co-contaminated
soil by Pleurotus cornucopiae and Bacillus thuringiensis FQ1 and the antioxidant responses in Pleurotus
cornucopiae. Ecotoxicol. Environ. Saf. 120: 386–393.
Jiang, L., C. Luo, D. Zhang, M. Song, Y. Sun and G. Zhang. 2018. Biphenyl-metabolizing microbial community and
a functional operon revealed in e-waste-contaminated soil. Environ. Sci. Technol. 52(15): 8558–8567.
Jørgensen, K. S. 2007. In situ bioremediation. Adv. Appl. Microbiol. 61: 285–305.
Juhasz, A. L. and R. Naidu. 2000. Bioremediation of high molecular weight polycyclic aromatic hydrocarbons: a
review of the microbial degradation of benzo [a] pyrene. Int. Biodeterior. Biodegradation. 45(1-2): 57–88.
Kamaludeen, S. P. B., K. R. Arunkumar and K. Ramasamy. 2003. Bioremediation of chromium contaminated
environments. Indian J. Exp. Biol. 41: 972–985.
Komancová, M., I. Jurčová, L. Kochánková and J. Burkhard. 2003. Metabolic pathways of polychlorinated biphenyls
degradation by Pseudomonas sp. 2. Chemosphere. 50(4): 537–543.
Kotoky, R. and P. Pandey. 2020. Rhizosphere assisted biodegradation of benzo (a) pyrene by cadmium resistant plant
probiotic Serratia marcescens S2I7, and its genomic traits. Sci. Rep. 10(1): 1–15.
Kumar, A., B. S. Bisht, V. D. Joshi and T. Dhewa. 2011. Review on bioremediation of polluted environment: a
management tool. Int. J. Environ. Sci. Technol. 1(6): 1079.
Kumar, V., S. K. Shahi and S. Singh. 2018. Bioremediation: an eco-sustainable approach for restoration of contaminated
sites. In Microbial bioprospecting for sustainable development (pp. 115–136). Springer, Singapore https://doi.
org/10.1007/978-981-13-0053-0_6.
16 Bioremediation for Sustainable Environmental Cleanup
Lade, H., S. Govindwar and D. Paul. 2015. Mineralization and detoxification of the carcinogenic Azo dye congo
red and real textile effluent by a polyurethane foam immobilized microbial consortium in an upflow column
bioreactor. Int. J. Environ. Res. Public Health. 12: 6894–6918.
Ladislas, S., A. El-Mufleh, C. Gérente, F. Chazarenc, Y. Andrès and B. Béchet. 2012. Potential of aquatic macrophytes
as bioindicators of heavy metal pollution in urban stormwater runoff. Water Air Soil Pollut. 223(2): 877–888.
Latha, A. P. and S. S Reddy. 2013. Review on bioremediation—Potential tool for removing environmental pollution.
Int. J. Basic Appl. Chem. Sci. 3(3): 21–33.
Lebeau, T., A. Braud and K. Jézéquel. 2008. Performance of bioaugmentation-assisted phytoextraction applied to
metal contaminated soils: a review. Environ. Pollut. 153(3): 497–522.
Lee, T. H., I. G. Byun, Y. O. Kim, I. S. Hwang and T. J. Park. 2006. Monitoring biodegradation of diesel fuel in
Bioventing processes using in situ respiration rate. Water Sci. Technol. 53(4-5): 263–272.
Lellis, B., C. Z. Fávaro-Polonio, J. A. Pamphile and J. C. Polonio. 2019. Effects of textile dyes on health and the
environment and bioremediation potential of living organisms. Biotechnol. Res. Innov. 3(2): 275–290.
Li, C. H., Y. S., Wong and N. F. Y. Tam. 2010. Anaerobic biodegradation of polycyclic aromatic hydrocarbons with
amendment of iron (III) in mangrove sediment slurry. Bioresour. Technol. 101(21): 8083–8092.
Li, Y., H. Gong, H. Cheng, L. Wang and M. Bao. 2017. Individually immobilized and surface-modified hydrocarbon-
degrading bacteria for oil emulsification and biodegradation. Mar. Pollut. Bull. 125(1-2): 433–439.
Li, Z., L. Wu, Y. Luo and P. Christie. 2018. Changes in metal mobility assessed by EDTA kinetic extraction in three
polluted soils after repeated phytoremediation using a cadmium/zinc hyperaccumulator. Chemosphere 194:
432–440.
Lin, H., C. Liu, B. Li and Y. Dong. 2021. Trifolium repens L. regulated phytoremediation of heavy metal contaminated soil
by promoting soil enzyme activities and beneficial rhizosphere associated microorganisms. J. Hazard. Mater.
402: 123829.
Lin, Q., X. Zhou, S. Zhang, J. Gao, M. Xie, L. Tao, F. Sun, M. Z. Hashmi and X. Su. 2022. Oxidative dehalogenation
and mineralization of polychlorinated biphenyls by a resuscitated strain Streptococcus sp. SPC0. Environ.
Res. 207: 112648.
Liu, J., H. Su, J. Xue and X. Wei. 2022. Optimization of decoloration conditions of methylene blue wastewater by
Penicillium P1. Indian J. Microbiol. 62(1): 103–111.
Maciel, B. M., A. C. F. Santos, J. C. T. Dias, R. O. Vidal, R. J. C. Dias, E. Gross and R. P. Rezende. 2009. Simple DNA
extraction protocol for a 16S rDNA study of bacterial diversity in tropical landfarm soil used for bioremediation
of oil waste. Genet. Mol. Res. 8(1): 375–388.
Maruthi, Y. A., K. Hossain and S. Thakre. 2013. Aspergillus flavus: A potential Bioremediator for oil contaminated
soils. Eur. J. Sustain. Dev. 2(1): 57–57.
Megharaj, M., B. Ramakrishnan, K. Venkateswarlu, N. Sethunathan and R. Naidu. 2011. Bioremediation approaches
for organic pollutants: a critical perspective. Environ. Int. 37(8): 1362–1375.
Mench, M., J. P. Schwitzguébel, P. Schroeder, V. Bert, S. Gawronski and S. Gupta. 2009. Assessment of successful
experiments and limitations of phytotechnologies: contaminant uptake, detoxification and sequestration, and
consequences for food safety. Environ. Sci. Pollut. Res.16(7): 876–900.
Motamedi, E., K. Kavousi, S. F. S. Motahar, M. R. Ghaffari, A. S. A. Mamaghani, G. H. Salekdeh and S. Ariaeenejad.
2021. Efficient removal of various textile dyes from wastewater by novel thermo-halotolerant laccase.
Bioresour. Technol. 337: 125468.
Murínová, S., K. Dercová and H. Dudášová. 2014. Degradation of polychlorinated biphenyls (PCBs) by four
bacterial isolates obtained from the PCB-contaminated soil and PCB-contaminated sediment. Int. Biodeterior.
Biodegradation. 91: 52–59.
Nayak, P. and H. Solanki. 2021. Pesticides and Indian agriculture—a review. Int. J. Res. Granthaalayah. 9: 250–263.
Nazifa, T. H., M. A. Ahmad, T. Hadibarata and A. Aris. 2018. Bioremediation of diesel oil spill by filamentous fungus
Trichoderma reesei H002 in aquatic environment. Int. J. Integr. Eng. 10(9).
Necibi, M. and N. Mzoughi. 2017. The distribution of organic and inorganic pollutants in marine
environment. Micropollutants: sources, Ecotoxicological Effects and Control Strategies. 129: 1–43.
Neilson, A. H. and A. S. Allard. 2008. Environmental degradation and transformation of organic chemicals, CRC
Press, Boca Raton, FL.
Omarova, M., L. T. Swientoniewski, I. K. M. Tsengam, A. Panchal, T. Yu, D. A. Blake, Y. M. Lvov and V. John. 2018.
Engineered clays as sustainable oil dispersants in the presence of model hydrocarbon degrading bacteria: the
role of bacterial sequestration and biofilm formation. ACS Sustain. Chem. Eng. 6(11): 14143–14153.
Ostrem Loss, E. M., M. K. Lee, M. Y. Wu, J. Martien, W. Chen, D. Amador-Noguez, C. Jefcoate, C. Remucal, S. Jung,
S. C. Kim and J. H. Yu. 2019. Cytochrome P450 monooxygenase-mediated metabolic utilization of benzo [a]
pyrene by Aspergillus species. MBio 10(3): e00558–19.
Pan, X., T. Xu, H. Xu, H. Fang and Y. Yu. 2017. Characterization and genome functional analysis of the DDT-
degrading bacterium Ochrobactrum sp. DDT-2. Sci. Total Environ. 592: 593–599.
Bioremediation: A Sustainable Approach for Environmental Cleanup 17
Pathak, A., A. Chauhan, A. Y. Ewida and P. Stothard. 2016. Whole genome sequence analysis of an Alachlor and
Endosulfan degrading Micrococcus sp. strain 2385 isolated from Ochlockonee River, Florida. Journal of
Genomics. 4: 42.
Prasad, M. N. V., P. Malec, A. Waloszek, M. Bojka and K. Strzallka. 2001. Physiological responses of Lemna trisulca L.
(duckweed) to cadmium and copper bioaccumulation. Plant Sci. 161: 881.
Priyanka, J. V., S. Rajalakshmi, P. S. Kumar, V. G. Krishnaswamy, D. A. Al Farraj, M. S. Elshikh and M. R. A.
Gawwad. 2022. Bioremediation of soil contaminated with toxic mixed reactive azo dyes by co-cultured cells
of Enterobacter cloacae and Bacillus subtilis. Environ. Res. 204: 112136.
Rezania, S., M. Ponraj, M. F. M. Din, A. R. Songip, F. M. Sairan and S. Chelliapan. 2015. The diverse applications
of water hyacinth with main focus on sustainable energy and production for new era: an overview. Renew.
Sustain. Energy Rev. 41: 943–954.
Rossi, T., P. M. S. Silva, L. F. De Moura, M. C. Araújo, J. O. Brito and H. S. Freeman. 2017. Waste from eucalyptus
wood steaming as a natural dye source for textile fibers. J. Clean. Prod. 143: 303–310.
Rudin, S. M., D. W. Murray and T. J. Whitfeld. 2017. Retrospective analysis of heavy metal contamination in Rhode
Island based on old and new herbarium specimens. Appl. Plant Sci. 5(1): 1600108.
Sagarkar, S., P. Bhardwaj, V. Storck, M. Devers-Lamrani, F. Martin-Laurent and A. Kapley. 2016. s-triazine degrading
bacterial isolate Arthrobacter sp. AK-YN10, a candidate for bioaugmentation of atrazine contaminated
soil. Appl. Microbial. Biotechnol. 100(2): 903–913.
Sangkharak, K., A. Choonut, T. Rakkan and P. Prasertsan. 2020. The degradation of phenanthrene, pyrene, and
fluoranthene and its conversion into medium-chain-length polyhydroxyalkanoate by novel polycyclic aromatic
hydrocarbon-degrading bacteria. Curr. Microbiol. 77(6): 897–909.
Saptakee, S. 2011. Journal on Bioremediation for Oil spills.www.buzzle.com/articles.
Sarang, B., K. Richa and C. Ram. 2013. Comparative study of bioremediation of hydrocarbon fuel. Int. J. Biotechnol.
Bioeng. Res. 4: 677–686.
Sarkar, D., R. Datta and R. Hannigan. 2011. Geochemical cycling of trace and rare earth elements in Lake Tanganyika
and its major tributaries. pp. 135–171. In: Elsevier [ed.]. Concepts and Applications in Environmental
Geochemistry, 135.
Sarkar, S., A. Banerjee, U. Halder, R. Biswas and R. Bandopadhyay. 2017. Degradation of synthetic Azo Dyes of
textile industry: a sustainable approach using microbial enzymes. Water Conserv. Sci. Eng. 24(2): 121–131.
Seeger, M., M. Hernández, V. Méndez, B. Ponce, M. Córdova and M. González. 2010. Bacterial degradation and
bioremediation of chlorinated herbicides and biphenyls. J. Plant. Nutr. Soil Sci. 10(3): 320–332.
Sengupta, K., M. T. Swain, P. G. Livingstone, D. E. Whitworth and P. Saha. 2019. Genome sequencing and comparative
transcriptomics provide a holistic view of 4-nitrophenol degradation and concurrent fatty acid catabolism by
Rhodococcus sp. strain BUPNP1. Front. Microbial. 9: 3209.
Shan, H., H. D. Kurtz Jr. and D. L. Freedman. 2010. Evaluation of strategies for anaerobic bioremediation of high
concentrations of halomethanes. Water Res. 44(5): 1317–1328.
Shah, V. and A. Daverey. 2020. Phytoremediation: A multidisciplinary approach to clean up heavy metal contaminated
soil. Environ. Technol. Innov. 18: 100774.
Sharma, R. and K. C. Yeh. 2020. The dual benefit of a dominant mutation in Arabidopsis IRON DEFICIENCY
TOLERANT1 for iron biofortification and heavy metal phytoremediation. Plant Biotechnol. J. 18(5): 1200–
1210.
Silva-Castro, G. A., I. Uad, J. Go´nzalez-Lo´pez, C. G. Fandino˜, F. L. Toledo and C. Calvo. 2012. Application of
selected microbial consortia combined with inorganic and oleophilic fertilizers to recuperate oil-polluted soil
using land farming technology. Clean Technol. Environ. Pol. 14: 719–726.
Singh, S. P. 2014. Application of bioremediation on solid waste management: a review. J. Bioremed. Biodegr. 5(06).
Singh, G. and S. K. Dwivedi. 2022. Biosorptive and biodegradative mechanistic approach for the decolorization of
congo red dye by Aspergillus species. Bull. Environ. Contam. Toxicol. 108(3): 457–467.
Smolyakov, B. S. 2012. Uptake of Zn, Cu, Pb, and Cd by water hyacinth in the initial stage of water system
remediation. Appl. Geochem. 27(6): 1214–1219.
Šrédlová, K., K. Šírová, T. Stella and T. Cajthaml. 2021. Degradation products of polychlorinated biphenyls and their
in vitro transformation by ligninolytic fungi. Toxics. 9(4): 81.
Syed, K., H. Doddapaneni, V. Subramanian, Y. W. Lam and J. S. Yadav. 2010. Genome-to-function characterization of
novel fungal P450 monooxygenases oxidizing polycyclic aromatic hydrocarbons (PAHs). Biochem. Biophys.
Res. Commun. 399(4): 492–497.
Tang, W. 2018. Research progress of microbial degradation of organophosphorus pesticides. Prog. Appl. Microbiol.
1: 29–35.
Thakor, R., H. Mistry, K. Tapodhan and H. Bariya. 2022. Efficient biodegradation of Congo red dye using fungal
consortium incorporated with Penicillium oxalicum and Aspergillus tubingensis. Folia Microbiol. 67(1): 33–43.
18 Bioremediation for Sustainable Environmental Cleanup
Tyagi, B. and N. Kumar. 2021. Bioremediation: principles and applications in environmental management.
In Bioremediation for Environmental Sustainability. 3–28. Elsevier.
Uday, U. S. P., T. K. Bandyopadhyay and B. Bhunia. 2016. Bioremediation and detoxification technology for treatment
of dye (s) from textile effluent. Textile Wastewater Treatment, 75–92.
United States Environmental Protection Agency (USEPA). 2000. Introduction Phytoremediation. EPA 600/R-99/107,
U.S. Environmental Protection Agency, Office of Research and Development, Cincinnati, OH.
Upendar, G., S. Dutta, P. Bhattacharya and A. Dutta. 2017. Bioremediation of methylene blue dye using Bacillus
subtilis MTCC 441. Water Sci. Technol. 75(7): 1572–1583.
U.S. EPA Seminars. Bioremediation of Hazardous Waste Sites: Practical Approach to Implementation, EPA/625/
K96/ 001.
Van Aken, B. 2009. Transgenic plants for enhanced phytoremediation of toxic explosives. Curr. Opin. Biotechnol. 20(2):
231–236.
Van Aken, B. and R. Bhalla. 2011. Microbial degradation of polychlorinated biphenyls. Ind. Toxic Wastes 152166.
Varjani, S. J. and V. N. Upasani. 2017. Crude oil degradation by Pseudomonas aeruginosa NCIM 5514: influence of
process parameters Indian J. Exp. Biol. 55: 493–497.
Varjani, S., P. Rakholiya, T. Shindhal, A. V. Shah and H. H. Ngo. 2021. Trends in dye industry effluent treatment and
recovery of value added products. J. Water Process. Eng. 39: 101734.
Vidali, M. 2001. Bioremediation. an overview. Pure Appl. Chem. 73(7): 1163–1172.
Vikrant, K., B. S. Giri, N. Raza, K. Roy, K. H. Kim, B. N. Rai and R. S. Singh. 2018. Recent advancements in
bioremediation of dye: current status and challenges. Bioresour. Technol. 253: 355–367.
Wagh, M. S., W. J. Osborne and S. Sivarajan. 2022. Bacillus xiamenensis and earthworm Eisenia fetida in bio removal
of lead, nickel and cadmium: a combined bioremediation approach. Appl. Soil Ecol. 176: 104459.
Wang, D. G., M. Yang, H. L. Jia, L. Zhou and Y. F. Li. 2009. Polycyclic aromatic hydrocarbons in urban street dust
and surface soil: comparisons of concentration, profile, and source. Arch. Environ. Contam. Toxicol. 56(2):
173–180.
Wang, X., Y. Wang, S. Ning, S. Shi and L. Tan. 2020. Improving azo dye decolorization performance and halotolerance
of Pichia occidentalis A2 by static magnetic field and possible mechanisms through comparative transcriptome
analysis. Front. Microbiol. 11: 1–10.
Whelan, M. J., F. Coulon, G. Hince, J. Rayner, R. McWatters, T. Spedding and I. Snape. 2015. Fate and transport of
petroleum hydrocarbons in engineered biopiles in polar regions. Chemosphere. 131: 232240.
Wiegel, J. and Q. Wu. 2000. Microbial reductive dehalogenation of polychlorinated biphenyls. FEMS Microb. Ecol.
32(1): 1–15.
Wolskm, E. A., V. Barrera, C. Castellari and J. F. González. 2012. Biodegradation of phenol in static cultures by
Penicillium chrysogenum ERK1: catalytic abilities and residual phototoxicity. Rev. Argent. Microbiol. 44(2):
113–121.
Yadav, J. S., H. Doddapaneni and V. Subramanian. 2006. P450ome of the white rot fungus Phanerochaete
chrysosporium: structure, evolution and regulation of expression of genomic P450 clusters.
Ye, Z., H. Li, Y. Jia, J. Fan, J. Wan, L. Guo, X. Su, Y Zhang, W. M. Wu and C. Shen. 2020. Supplementing resuscitation-
promoting factor (Rpf) enhanced biodegradation of polychlorinated biphenyls (PCBs) by Rhodococcus
biphenylivorans strain TG9T. Environ. Pollut. 263: 114488.
Yu, X. Z. and J. D. Gu. 2007a. Accumulation and distribution of trivalent chromium and effects on hybrid willow
(Salix matsudana Koidz × alba L.) metabolism. Arch. Environ. Contam. Toxicol. 52: 503–511.
Yu, X. Z. and J.D. Gu. 2007b. Metabolic responses of weeping willows to selenate and selenite. Env. Sci. Pollut. Res.
14: 510–517.
Yu, X. Z. and J. D. Gu. 2008a. The role of EDTA in phytoextraction of hexavalent and trivalent chromium by two
willow trees. Ecotoxicol. 17:143-152.
Yu, X. Z. and J. D. Gu. 2008b. Effect of available nitrogen on phytoavailability and bioaccumulation of hexavalent
and trivalent chromium in hankow willows (Salix matsudana Koidz) Ecotoxicol. Environ. Saf. 70: 216–222.
Zhang, J. L. and C. L. Qiao. 2002. Novel approaches for remediation of pesticide pollutants. Int. J. Environ.
Pollut. 18(5): 423–433.
Zhang, W., Z. Lin, S. Pang, P. Bhatt and S. Chen. 2020. Insights into the biodegradation of lindane
(γ-hexachlorocyclohexane) using a microbial system. Frontiers in Microbiology. 11: 522.
Zheng, M., W. Wang, M. Hayes, A. Nydell, M. A. Tarr, S. A. Van Bael and K. Papadopoulos. 2018. Degradation of
Macondo 252 oil by endophytic Pseudomonas putida. J. Environ. Chem. Eng. 6(1): 643–648.
Zhou, H., S. Zhang, J. Xie, H. Wei, Z. Hu and H. Wang. 2020. Pyrene biodegradation and its potential pathway
involving Roseobacter clade bacteria. Int. Biodeterior. Biodegrad. 150: 104961.
Zhu, D. H., F. H. Nie, Q. L. Song, W. Wei, M. Zhang, Y. Hu, H. Y. Lin, D. J. Kang, Z. B. Chen and J. J. Chen. 2022.
Isolation and genomic characterization of Klebsiella Lw3 with polychlorinated biphenyl degradability. Environ.
Technol., 1–11.
Chapter 2
Bioprecipitation as a
Bioremediation Strategy for
Environmental Cleanup
Samantha M. Wilcox,1 Catherine N. Mulligan1,* and
Carmen Mihaela Neculita2
2.1 Introduction
The term ‘sustainable’ is generally used in today’s culture and its true significance is often not clear.
While typically used synonymously to environmentally friendly behavior, the term represents much
more. According to the United Nations (UN) 17 Sustainable Development Goals, ‘sustainability’
refers to social, economic and environmental action (UNDP 2022). The term ‘sustainability’, used
throughout this chapter, aligns itself with the United Nations definition.
Imperative to the achievement of ‘sustainability’ is its implementation during decision-making
processes. The introduction of ISO Lifecycle Assessment 14040 and the field of Environmental
Accounting has made it easier to adhere to sustainable practices in government and industry (ICF
Incorporated 1995, ISO 2006). These approaches aim to minimize the environmental, social and
economic impact at all stages of a product, a process or a service.
Remediation strategies, especially bio-based processes, are often considered inherently
sustainable. However, it is important to apply these principles to improve longstanding engineering
frameworks. An effective engineered clean-up strategy will appease the sustainability requirements
and offer solutions that meet both environmental and governmental regulations. This chapter focuses
on biological precipitation, also referred to as bioprecipitation as a remediation strategy for soil and
groundwater contamination. The process enhances the already naturally occurring processes. It is
therefore considered a cost effective, socially accepted and environmentally friendly technique.
This chapter aims to document and justify the necessity and highlight the sustainable nature of
bioprecipitation as a technique for environmental clean-up.
1
Department of Building, Civil and Environmental Engineering, Concordia University, Montreal, Canada.
2
Research Institute on Mines and the Environment (RIME), University of Quebec in Abitibi-Témiscamingue, Rouyn-
Noranda, Canada.
* Corresponding author: [email protected]
20 Bioremediation for Sustainable Environmental Cleanup
Figure 2.1. Human exposure to metal(loid) contamination including contact type, exposure pathways, exposure routes, and
impacted organs (Adapted from Kuppusamy et al. 2020).
Table 2.1. Clean-up strategies (Adapted from Government of Canada 2017, LaGrega et al. 1994, Ossai et al. 2020).
2.3 Bioprecipitation
2.3.1 Overview: Principles and Applications
2.3.1.1 Chemical Precipitation
Precipitation is an intricate phenomenon involving thermodynamic and kinetic processes. The
process is governed by its thermodynamic properties, i.e., supersaturation state. A solution is
considered supersaturated when the solute and solvent are no longer in an equilibrium (saturated)
state (Karpiński and Bałdyga 2019, Lewis 2017). Equation 2.1 shows the supersaturation calculation,
whereby the supersaturation (σ) is calculated from the actual molar chemical potential (μ), the
molecule in equilibrium state (μeq), the universal gas constant (R) and the absolute temperature (T).
Based on the equation:
- The solution is in equilibrium when ∆μ (∆μ = μ – μeq) is equal to 0
- If ∆μ > 0 the solution is supersaturated, spontaneous precipitation will occur
- If ∆μ < 0 the solution is below the saturated state, spontaneous dissolution will occur (Karpiński
and Bałdyga 2019).
Supersaturation (Davey and Garside 2000, Karpiński and Bałdyga 2019)
(μ – μeq)
σ= Eq. 2.1
RT
Phase diagrams can demonstrate the liquid or solid phase of a compound of interest (Karpiński
and Bałdyga 2019). The thermodynamic component of precipitation can be better understood by the
Gibbs phase rule, which identifies the number of possible phases and the degree of freedom of the
multiphase system in equilibrium (Faghri and Zhang 2006). The Gibbs phase rule both identifies
and inhibits solid precipitates formed from a solution (Yong et al. 2014).
Furthermore, the thermodynamic potential is described by Gibbs free energy. It explains the
maximum energy transfer of a closed system. The change in Gibbs free energy (∆G) can be used
to define primary homogenous nucleation. Equation 2.2 shows the calculation, where the number
of molecules (N), the reaction affinity (Φ), the crystal surface area (A) and the surface tension (σ)
are used to explain the work to produce crystals during precipitation of a supersaturated solution.
From Eq. 2.1, one can consider R as the molar chemical potential change or the Gibbs free energy
(Karpiński and Bałdyga 2019).
Gibbs Free Energy Change (Karpiński and Bałdyga 2019, Nielsen 1964)
While supersaturation state is the driving force of precipitation, the kinetic process illustrates
the rate of precipitation. Nucleation, crystal growth and agglomeration represent the primary kinetic
process involved in precipitation (Lewis 2017), these terms are further defined in Table 2.2. Again
∆μ
from Eq. 2.1, the term represents the rate of nucleation and crystal growth (Karpiński and
Bałdyga 2019). RT
Precipitation can be applied to soil, groundwater, surface water and wastewater treatment. For
soil and groundwater remediation, metal(loid)s can chemically precipitate out of the pore water
(Yong et al. 2014). These precipitates can adsorb onto soil particle surfaces (Yong et al. 2014) or can
form a cement matrix clogging the pore spaces (Mitchell and Soga 2005). For water and wastewater
24 Bioremediation for Sustainable Environmental Cleanup
treatment, chemical precipitation is applied through coagulation and flocculation, whereby the use
of a reagent is used to promote precipitation and aggregation, respectively (Davis 2010, Mihelcic
and Zimmerman 2014). Figure 2.2 demonstrates the various factors that affect precipitation and
potential impacts to its use as a remediation strategy. Of the factors identified, pH is the most
significant, whereby alkaline solutions tend to precipitate metal(loid)s and acidic solutions cause
dissolution of metal(loid) precipitates (Yong et al. 2014).
and
Figure 2.2. Factors influencing precipitation (Adapted from Yong et al. 2014).
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 25
Metal Precipitation from Hydrogen Sulfide (Sahinkaya et al. 2017, Willis and Donati 2017)
SRB species can be both heterotrophs and autotrophs (chemolithotrophs) found in anaerobic
environments (Barton et al. 2015, Hao et al. 2014). As of 2015, there were 59 genera and
220 species reported (Barton et al. 2015). However, the genera Desulfovibrio is the most highly
reported bacteria from bioreactor studies (Kiran et al. 2017). The microorganisms are classified
as complete oxidizers, incomplete oxidizers or both (Hao et al. 2014). There are over 75 energy
sources that promote SRB growth (Barton et al. 2015). Although trace metals, selenium (Se)
and molybdenum (Mo), are attributed to bacterial growth, the mechanisms linked to metal(loid)
reduction are not currently considered as related (Barton et al. 2003).
To assure microorganisms can facilitate bioprecipitation, they require some degree of metal
tolerance. The metal tolerance allows the microorganisms to flourish under normally toxic metal(loid)
contamination. This is achieved via metabolism, cell wall structure, Extra-cellular Polymeric
Substances (EPS), methylation, alkylation/dealkylation (Diels et al. 2006), intra-cellular and extra-
cellular sequestration, active transport effluent pumps, enzymatic detoxification, reduction in metal
sensitivity of cellular targets (Bruins et al. 2000, Diels et al. 2006) and exclusion by permeability
barrier (Bruins et al. 2000). Metal resistant genes have been found in both gram-positive and gram-
negative microorganisms (Abou-Shanab et al. 2007) and SRB species have also reported gram
26 Bioremediation for Sustainable Environmental Cleanup
positive and gram-negative microorganisms (Barton et al. 2015). To improve culture resistance a
mixture of species is often selected for bioprecipitation processes (Kiran et al. 2017).
In addition to BSR, SRB have also been found to reduce other metal(loid)s with direct or
indirect microorganism activity. SRB can directly reduce metal(loid)s (i.e., Cr, As, Al, Te and Sb)
to a less toxic, insoluble form or can indirectly reduce metal(loid)s via hydrogen sulfide produced
during BSR (Sánchez-Andrea et al. 2016, Willis and Donati 2017). After metal(loid) reduction,
many species are capable of precipitation. For example, uranium (U6+) can precipitate uraninite
(UO2), chromium (Cr6+) in the presence of ferric iron (Fe3+) can precipitate chromium hydroxide
oxide (CrO (OH)) and chromium sulfide (Cr2S3) and arsenic can precipitate arsenic sulfide (As2S3)
and arsenopyrite (FeAsS). Metal(loid) reduction using SRB activity is presented in Table 2.3. The
following electron donors were used as an energy source in some of the noted reactions: acetate
(CH3COOH), formate (CH2O2), lactate (C3H6O3) and pyruvate (C3H4O3).
Table 2.3. Reduction of Metal(loid)s with SRB (Adapted from/Barton et al. 2015, Lovley 1993, Sahinkaya et al. 2017).
Electron donors provide a carbon and energy source for the reaction. The selection of
an appropriate electron donor is the ratio of Chemical Oxygen Demand (COD) and sulfate ion
concentration (SO42–), i.e., COD/SO42– is important. There is a correlation between the interaction
of SRB with carbon source and electron acceptor (Barbosa et al. 2014, Kiran et al. 2017). The COD
denotes oxygen content required to oxidize the organic material and the electron acceptor is SO42–
(Barbosa et al. 2014). However, COD is measured under aerobic conditions, and therefore does not
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 27
accurately portray the organic carbon available to anerobic microorganisms (Neculita and Zagury
2008). The minimum theoretical value acceptable for organic degradation and BSR is 0.67 (Hao
et al. 1996, Kiran et al. 2017, Neculita and Zagury 2008).
Other noteworthy indicators for an electron donors’ performance are the carbon (C) to nitrogen
(N) ratio (C/N) and the Dissolved Organic Carbon (DOC) to SO42– ratio (DOC/SO42–) (Neculita
and Zagury 2008). The C/N ratio gives information about the capacity of biological degradation of
an electron donor (Reinertsen et al. 1984, Zagury et al. 2006). While the ratio does not necessarily
indicate the actual C and N available to the microorganisms (Reinertsen et al. 1984), a value of about
10 is generally accepted as a suitable substrate (Béchard et al. 1994, Neculita and Zagury 2008,
Reinertsen et al. 1984, Zagury et al. 2006). The DOC/SO42– ratio is like the COD/SO42– ratio but is
more easily quantified (Neculita and Zagury 2008).
An electron donor can be organic or synthetic. In keeping with the sustainable nature of
remediation, an organic energy is a natural, economic and socially acceptable additive to the
operation. Organic waste can be considered sewage sludge, animal manure, leaf mulch, wood chips,
sawdust and cellulose (Liamleam and Annachhatre 2007). The high carbon content of organic waste
is advantageous to BSR demonstrating high rates of sulfate reduction, specifically when wastes
are applied as mixtures (Hao et al. 2014, Liamleam and Annachhatre 2007). The COD/SO42– ratio
for organic wastes ranges from 1.6–5 (Kiran et al. 2017). In general, for all performance indicators
(i.e., C/N, COD/SO42–, DOC/SO42–), a higher ratio is linked to superior BSR (Neculita and Zagury
2008).
Molasses is the most cost effective and widely available electron donor for BSR (Janssen and
Temminghoff 2004, Liamleam and Annachhatre 2007), indicating its preferential use for sustainable
engineering. In addition to its use as an electron donor, it contains nutrients (i.e., P, K, Cl, amino
acids) for SRB growth (Janssen and Temminghoff 2004). The process aims to ferment molasses into
lactate which is used as the electron donor or carbon source (Liamleam and Annachhatre 2007).
However, SRB growth is inhibited at molasses concentrations greater than 5 g/L (Hao et al. 2014,
Janssen and Temminghoff 2004), and there is seemingly a link between the quantity applied and
the pH rise (Janssen and Temminghoff 2004). Further, partial decomposition creates high COD in
effluent (Hao et al. 2014).
Lactate is another organic substrate used as an energy source for BSR. Its use as an electron donor
is not temperature dependent, and does not impact its oxidizing ability (Liamleam and Annachhatre
2007). Both lactate and ethanol are considered optimal for SRB growth (Janssen and Temminghoff
2004), although, ethanol is not an organic substrate it is considered the most cost effective (Gibert
et al. 2002, Liamleam and Annachhatre 2007). Further, acetate is the most used energy source for
Fe3+ bioprecipitation (Lovley 1993). Other energy sources include hydrogen, formate, methanol,
propionate, butyrate, sugar and hydrocarbons (Liamleam and Annachhatre 2007). Zero-valent iron
has also been used to establish an anerobic environment for SRB growth by consuming oxygen
(O2) while generating hydrogen (H2) which acts as an electron donor (Pagnanelli et al. 2009). A
high value for Gibbs free energy is preferable for BSR to assure that sulfidogenic reactions prevail
over methanogenic reactions (Liamleam and Annachhatre 2007). However, the concentration of
the electron donor should be monitored as high carbon concentration can lead to methanogenic
conditions (Diels et al. 2005, 2006). A mixture of organic substrates is often used with success,
that can produce high rates of sulfate reduction (Liamleam and Annachhatre 2007). The cost of the
substrate should be kept in mind and adhere to the projects’ economic sustainability.
Bioprecipitation can also occur as an oxidative reaction. Ferrous iron (Fe2+) is transformed via
microbial oxidation into ferric iron (Fe3+) for precipitation. Microbial oxidation can take place as
an anaerobic process (using phototropic and nitrate-reducing microorganisms) and aerobic process
(using neutrophilic and acidophilic microorganisms). The phylum Proteobacteria is the most common
bacteria responsible for Fe2+ reduction (Kiskira et al. 2017). Zero-valent iron can oxidize to Fe2+ and
Fe3+ ions, help to facilitate the bioprecipitation process (Pagnanelli et al. 2009). From the oxidative
reaction, Fe3+ can precipitate ferric hydroxide (Fe (OH)3), jarosite (MFe3(SO4)2(OH)6, where M is a
28 Bioremediation for Sustainable Environmental Cleanup
monovalent cation, i.e., K+, Na+, NH4+, Ag+, or H3O+), schwertmannite (Fe16O16(SO4)2(OH)6 • nH2O,
where n is between 10 and 12), goethite (FeO (OH)), hematite (Fe2O3) and scorodite (FeAsO4 • 2H2O).
The pH, temperature, and overall chemistry of solution influence the precipitate formed (Sahinkaya
et al. 2017).
The aim of this process is to immobilize the precipitates through the formation of a cement
matrix, whereby precipitates form bridges between soil particles. Since the crystals precipitate
out of the soil-groundwater system and clog the pore spaces, the soil properties are altered. The
permeability, porosity, stiffness, shear strength, unconfined compressive strength, microstructure
and shear wave velocity are all impacted by MICP (Mujah et al. 2016).
The main theory behind MICP remediation is solidification/stabilization (S/S). S/S is a strategy
that immobilizes the soil and groundwater contaminants by using additives that alters the physical
properties (i.e., solidifies/entraps the contaminant) and/or chemical properties (i.e., transforms
the contaminant to a less toxic, less mobile form), respectively (LaGrega et al. 1994, Sharma
and Reddy 2004). Historically, S/S used cements, pozzolans, thermoplastic materials or organic
polymers to achieve contaminant entrapment (Sharma and Reddy 2004), however MICP offers a
biological approach to reach S/S remediation. The mechanisms involved in S/S to remediate soil and
groundwater include macroencapsulation, microencapsulation, adsorption, absorption, precipitation
and detoxification (LaGrega et al. 1994). Through MICP, sorption can cause bioprecipitated
CaCO3 crystals and other MCO3 compounds to bond to soil particle surfaces (LaGrega et al. 1994,
Xiangliang 2009) via electrochemical bonds, such as van der Waal’s forces or hydrogen bonds
(LaGrega et al. 1994). This can aid the development of the cement matrix, which offers a more
sustainable approach to S/S remediation.
The most common microorganism used for MICP is the Bacillus species (Achal and Pan 2011).
To improve MICP performance a desirable microorganism should be high urease producing with
high metal tolerance. Again, these organisms can be gram-positive or gram-negative, however
gram-positive bacteria are more reactive (Beveridge and Fyfe 1985, Levett et al. 2020). The
negative charge of the bacterial cell wall can attract Ca2+ ions causing MICP on the gram-positive
and gram-negative cell walls (Achal and Pan 2011). Microorganism EPS can also affect MICP.
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 29
EPS is reported to impact the biofilm, cell adhesion, CaCO3 capture (Achal and Pan 2011) and
the CaCO3 mineralogy (Kawaguchi and Decho 2002). The CaCO3 mineralogy can be altered via
various polymorphs (i.e., calcite, vaterite and aragonite). The type of polymorph produced during
MICP can impact the stability of precipitates, where calcite is the most stable and desirable, while
aragonite and vaterite are less stable.
Other factors influencing the efficacy of MICP as a remediation technique are temperature,
bacterial concentration or density, pH, degree of saturation, concentration of cementation solution
and field application (Mujah et al. 2016). This bioprecipitation method offers a promising remediation
strategy. However, the long-term impact should be studied. Metal(loid) dissolution from redox and/
or pH changes could release contaminants into the soil and groundwater. Additionally, over time
metal(loid) leaching can occur from cracks, fissures or interstices formed in the cement matrix.
These defects can be caused by wind, erosion, wetting and drying cycles, hot and cold cycles, rain
and snow and/or other environmental elements.
2.3.2 Design
Bioprecipitation can be applied as a remediation solution in a variety of methods. The processes are
classified as in-situ or ex-situ based on where the operation takes place. If clean-up occurs at the
site, the operation is considered an in-situ process. However, if soil is extracted from the site and
transported for treatment, the operation is an ex-situ process. In addition to these operation models,
bioprecipitation often occurs either simultaneously or in sequence with other treatment methods.
This is based on the requirements of the project, i.e., future land usage and the desired level of
metal(loid) removal. This chapter will put emphasis on in-situ bioremediation strategies that are
passive or rely on natural attenuation processes. These strategies are innately more sustainable and
therefore, offer benefits as a clean-up protocol.
metal(loid) contaminated plume, whereby the reactive barrier is designed to degrade or immobilize
the contaminant via BSR. The two primary configurations of a reactive barrier are continuous
(vertical barrier, perpendicular to the contaminant plume) and funnel-and-gate system (V-shaped
funnel directing contaminant plume through the vertical reactive gate) (Sharma and Reddy 2004).
The reactive barrier is designed for the specific site, such that reactive material (electron donor and
microorganism consortium) is selected based on the desired type of bioprecipitation. Acid Mine
Drainage (AMD) for example, which is highly acidic and heavily contaminated with sulfuric acid
(SO42– and H+) and other heavy metals, could use a layered mixture of silica sand, organic waste
and silica sand either in a horizontal or vertical sequence to decrease effluent AMD (Benner et al.
1999, Waybrant et al. 2002). Clogging of reactive material due to bioprecipitation may decrease the
efficacy of the barrier over time (Kiran et al. 2017).
Wetlands and engineered wetlands aim to remove metal(loid) contaminants from water with
degradation and bioprecipitation techniques. They can be in the form of aerobic wetlands, anaerobic
wetlands (Johnson and Santos 2020) or anoxic ponds (Kiran et al. 2017). There are two main
impacts to the redox potential of these designs: the hydraulic design and the mode of operation. The
hydraulic designs can use a vertical flow (aerobic) treatment, horizontal subsurface flow (anoxic)
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 31
treatment, subsurface flow treatment (anoxic) (Faulwetter et al. 2009) or a surface flow (aerobic)
treatment (Kosolapov et al. 2004). The mode of operation refers to the feed mode (batch feed,
intermittent flow feed or continuous flow feed), hydraulic load rate and Hydraulic Retention Time
(HRT) (Faulwetter et al. 2009). The use of plant species to help mitigate pollution via the influence
of the redox condition is specific to wetlands (Faulwetter et al. 2009) and phytoremediation (Stephen
and Macnaughtont 1999). Phytoremediation can mitigate soil and groundwater pollution via five
primary mechanisms: phytostabilization, phytoextraction, phytovolatilization, rhizodegradation and
phytotransformation. Rhizodegradation, specifically, is applicable during wetland treatment, as the
plant-soil-microorganism system in the rhizosphere can degrade contaminants (Shmaefsky 2020).
Biofilters offer an additional method to remediate soil and groundwater via in-situ
bioprecipitation. The method consists of a filter media that allows microorganisms to attach and
multiply. The biofilter essentially acts as a surface for microorganism immobilization leading to
the development of a biofilm, which enables redox reactions necessary for bioprecipitation. There
are numerous organic and inorganic filter materials that can be used to facilitate biofiltration. In
keeping with the sustainable assessment of clean-up strategies, compost can be used as a viable,
environmental filter media. The compost media has nutrients for microorganism growth, good
water retention capacity for microorganism metabolism and good permeability for homogenous
distribution that are beneficial to the biofilter process (Pachaiappan et al. 2022).
MICP is typically applied to the field via injection, surface percolation or pre-mixing. The latter
two applications are in-situ treatment methods and therefore preferential. An issue experienced
with MICP remediation is the uniform distribution of the mixture (cementation solution and
microorganism) to adequately precipitate a CaCO3 matrix. An injection can lead to clogging around
the injection site, which inhibits its performance as a clean-up strategy. By reducing the injection rate,
the mixture may increase its reach. Surface percolation, however, has a better uniform distribution.
The mixture percolates through the soil via gravity. The depth of the required remediation should
be previously assessed as the permeability of soil impacts the depth reached by percolation (Mujah
et al. 2016).
Sorption and bioprecipitation are remediation mechanisms that typically occur simultaneously.
Sorption includes adsorption (accumulation to surfaces) or absorption (penetration into substances).
As mentioned earlier, sorption is part of the S/S technique aiming to immobilize metal(loid)
contaminants in-situ (LaGrega et al. 1994). Once contaminants precipitate from groundwater or pore
space, adsorption to soil particle surfaces is desirable. Again, considering sustainable remediation,
the use of an organic matter energy source during bioprecipitation can enhance the sorption capacity.
The carboxylic group of olive pomace, compost and leaves can facilitate sorption (Pagnanelli et al.
2009), improving immobilization.
These in-situ remediation methods offer a greener solution for industrial clean-up of soil and
groundwater. Although some greenhouse gas emissions and energy will be consumed to either drill
the injection wells or implement the reactive barrier, these strategies offer better sustainability to
ex-situ methods. However, it is important to note that these strategies are not viable for all types of
contamination. Often more intensive methods are required to achieve high degrees of metal(loid)
contaminant removal required by environmental and government agencies.
An overview of reactor designs for ex-situ bioprecipitation is described in Table 2.5. In addition
to the design itself, a reactor can operate under different modes or various stages. The operation
mode refers to how the influent is added to the reactor, i.e., batch, continuous or semi-continuous
application. The stages refer to the sequence of operations. In a one-stage process bioprecipitation
occurs in one reactor, while a two-stage process will separate mechanisms (i.e., oxidative-reductive
processes and precipitation) to occur in different reactors (Sánchez-Andrea et al. 2014).
Land Use
and and
and
and
Figure 2.3. Selection process for clean-up strategy incorporating site assessment, land use requirements, feasibility and
implementation (Adapted from Wilcox et al. 2022).
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 33
Table 2.5. Ex-situ reactor designs for bioprecipitation (Adapted from Kaksonen and Puhakka 2007, Kiran et al. 2017,
Speece 1983).
Downflow Advantages:
Anaerobic Filter - Utilizes gravitation force
Reactor (DAFR) - Good sludge retention
- Low shear force
Disadvantages:
- Channeling due to shear force
- Large pressure gradients
Fluidized-Bed Advantages:
Reactor (FBR) - Large surface area for biofilm
- Good biomass retention
- Small pressure gradients
- No channeling or clogging
- Large mass transfer rates
Disadvantages:
- Biomass loss due to shear force
- Energy requirement
Membrane Advantages:
Bioreactor (MBR) - Good biomass retention
Disadvantages:
- Fouling
* Red arrow represents gas, blue arrow represents influent and effluent, and green arrow represents recycle process.
2.4 Conclusions
Bioprecipitation offers a sustainable approach to soil and groundwater clean-up for metal(loid)
contamination. The process offers social, economic and environmental benefits as a sustainable
remediation strategy. As a biological treatment method using natural processes, public acceptance
tends to be high in comparison to a more invasive treatment, such as physical, chemical, thermal or
electrical operations.
While bioprecipitation can occur as an in-situ and ex-situ remediation technique, in-situ
treatment offers a more sustainable approach. The operation occurs at the site with minimal soil
disturbance, energy consumption and greenhouse gas emissions. The economic value is also high as
capital, maintenance and transport costs are relatively low.
Bioprecipitation can occur as an oxidative-reductive reaction or a cementation technique. The
desired mechanism of immobilization should be based on the type and degree of contamination
with a focus on chemical stability. For oxidative-reductive reactions an energy source should be
selected to facilitate the reaction. Organic waste is a viable electron donor, showing promise as a
carbon source with high COD/SO42– ratios for BSR. Organic waste can be repurposed to a valuable
bioprecipitation additive, enhancing the sustainable nature of the operation.
A thorough site assessment should be conducted at each site prior to implementation as a
remediation strategy. While bioprecipitation shows promise as a bioremediation strategy for
metal(loid) clean-up, laboratory testing is required to better predict its applicability for each
individual case . The strategy should be selected only if it meets the project objectives, the future land
use requirements and can achieve high removal efficacies that adhere to environmental standards.
36 Bioremediation for Sustainable Environmental Cleanup
References
Abou-Shanab, R. A. I., P. van Berkum and J. S. Angle. 2007. Heavy metal resistance and genotypic analysis of
metal resistance genes in gram-positive and gram-negative bacteria present in Ni-rich serpentine soil
and in the rhizosphere of Alyssum murale. Chemosphere. 68(2): 360–367. https://doi.org/10.1016/j.
chemosphere.2006.12.051.
Achal, V. and X. Pan. 2011. Characterization of urease and carbonic anhydrase producing bacteria and their role
in calcite precipitation. Current Microbiol. 62(3): 894–902. http://dx.doi.org.qe2a-proxy.mun.ca/10.1007/
s00284-010-9801-4.
Achal, V., X. Pan and D. Zhang. 2011. Remediation of copper-contaminated soil by Kocuria flava CR1, based
on microbially induced calcite precipitation. Ecol. Eng. 37(10): 1601–1605. https://doi.org/10.1016/j.
ecoleng.2011.06.008.
Achal, V., A. Mukerjee and M. Sudhakara Reddy. 2013a. Biogenic treatment improves the durability and remediates the
cracks of concrete structures. Constr. Build. Mater. 48: 1–5. https://doi.org/10.1016/j.conbuildmat.2013.06.061.
Achal, V., X. Pan, D.-J. Lee, D. Kumari and D. Zhang. 2013b. Remediation of Cr(VI) from chromium slag by
biocementation. Chemosphere. 93(7): 1352–1358. https://doi.org/10.1016/j.chemosphere.2013.08.008.
Barbosa, L. P., P. F. Costa, S. M. Bertolino, J. C. C. Silva, R. Guerra-Sá, V. A. Leão and M. C. Teixeira. 2014. Nickel,
manganese and copper removal by a mixed consortium of sulfate reducing bacteria at a high COD/sulfate
ratio. World J. Microbiol. Biotechnol. 30(8): 2171–2180. https://doi.org/10.1007/s11274-013-1592-x.
Barton, L. L., R. M. Plunkett and B. M. Thomson. 2003. Reduction of metals and nonessential elements by anaerobes.
pp. 220–234. In: L. G. Ljungdahl, M. W. Adams, L. L. Barton, J. G. Ferry and M. K. Johnson [eds.].
Biochemistry and Physiology of Anaerobic Bacteria. Springer. https://doi.org/10.1007/0-387-22731-8_16.
Barton, L. L., F. A. Tomei-Torres, H. Xu and T. Zocco. 2015. Metabolism of metals and metalloids by the sulfate-
reducing bacteria. pp. 57–83. In: D. Saffarini [ed.]. Bacteria-Metal Interactions. Springer International
Publishing. https://doi.org/10.1007/978-3-319-18570-5_4.
Béchard, G., H. Yamazaki, W. D. Gould and P. Bédard. 1994. Use of cellulosic substrates for the microbial treatment of acid
mine drainage. J. Environ. Qual. 23(1): 111–116. https://doi.org/10.2134/jeq1994.00472425002300010017x.
Benner, S. G., D. W. Blowes, W. D. Gould, R. B. Herbert and C. J. Ptacek. 1999. Geochemistry of a permeable
reactive barrier for metals and acid mine drainage. Environ. Sci. Technol. 33(16): 2793–2799. https://doi.
org/10.1021/es981040u.
Beveridge, T. J. and W. S. Fyfe. 1985. Metal fixation by bacterial cell walls. Can. J. Earth Sci. 22(12): 1893–1898.
https://doi.org/10.1139/e85-204.
Bruins, M. R., S. Kapil and F. W. Oehme. 2000. Microbial resistance to metals in the environment. Ecotoxicol.
Environ. Saf. 45(3): 198–207. https://doi.org/10.1006/eesa.1999.1860.
Davey, R. J. and J. Garside. 2000. From Molecules to Crystallizers: An Introduction to Crystallization. Oxford
University Press.
Davis, M. L. 2010. Water and Wastewater Engineering: Design Principles and Practice. The McGraw-Hill
Companies, Inc.
Diels, L., J. Geets, W. Dejonghe, S. van Roy, K. Vanbroekhoven and G. Malina. 2005. Heavy metal immobilization
in groundwater by in situ bioprecipitation: comments and questions about carbon source use, efficiency and
sustainability of the process. Proceedings of the 9th International FKZ/TNO Conference on Contaminated Soil
(Consoil). France, 355–360.
Diels, L., J. Geets, W. Dejonghe, S. V. Roy, K. Vanbroekhoven, A. Szewczyk and G. Malina. 2006. Heavy metal
immobilization in groundwater by in situ bioprecipitation: comments and questions about efficiency and
sustainability of the process. Proceedings of the Annual International Conference on Soils, Sediments, Water
and Energy. Amherst. 11: 99–112. https://scholarworks.umass.edu/soilsproceedings/vol11/iss1/7.
Faghri, A. and Y. Zhang. 2006. Chapter 2: thermodynamics of multiphase systems. pp. 107–176. In: Transport
Phenomena in Multiphase Systems. Elsevier Academic Press.
Faulwetter, J. L., V. Gagnon, C. Sundberg, F. Chazarenc, M. D. Burr, J. Brisson, A. K. Camper and O. R. Stein. 2009.
Microbial processes influencing performance of treatment wetlands: a review. Ecol. Eng. 35(6): 987–1004.
https://doi.org/10.1016/j.ecoleng.2008.12.030.
Gadd, G. M. 2004. Microbial influence on metal mobility and application for bioremediation. Geoderma. 122(2):
109–119. https://doi.org/10.1016/j.geoderma.2004.01.002.
Gibert, O., J. de Pablo, J. L. Cortina and C. Ayora. 2002. Treatment of acid mine drainage by sulphate-reducing
bacteria using permeable reactive barriers: a review from laboratory to full-scale experiments. Rev. Environ.
Sci. Biotechnol. 1(4): 327–333. https://doi.org/10.1023/A:1023227616422.
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 37
Government of Canada. 2017. Compare decontamination technologies—Guidance and Orientation for the Selection
of Technologies—Contaminated sites—Pollution and waste management—Environment and natural
resources—Canada.ca. https://gost.tpsgc-pwgsc.gc.ca/Techlst.aspx?lang=eng#wb-auto-5.
Hao, O. J., J. M. Chen, L. Huang and R. L. Buglass. 1996. Sulfate‐reducing bacteria. Crit. Rev. Environ. Sci. Technol.
26(2): 155–187. https://doi.org/10.1080/10643389609388489.
Hao, T., P. Xiang, H. R. Mackey, K. Chi, H. Lu, H. Chui, M. C. M. van Loosdrecht and G.-H. Chen. 2014. A review
of biological sulfate conversions in wastewater treatment. Water Res. 65: 1–21. https://doi.org/10.1016/j.
watres.2014.06.043.
Hengen, T. J., M. K. Squillace, A. D. O’Sullivan and J. J. Stone. 2014. Life cycle assessment analysis of active
and passive acid mine drainage treatment technologies. Resour. Conserv. Recycl. 86: 160–167. https://doi.
org/10.1016/j.resconrec.2014.01.003.
ICF Incorporated. 1995. An introduction to environmental accounting as a business management tool: key concepts
and terms (EPA 742-R-95-001). https://www.epa.gov/sites/default/files/2014-01/documents/busmgt.pdf.
ISO. 2006. ISO 14040:2006(en), environmental management—life cycle assessment—principles and framework.
International Organization for Standardization. https://www.iso.org/obp/ui/#iso:std:iso:14040:ed-2:v1:en.
Janssen, G. M. C. M. and E. J. M. Temminghoff. 2004. In situ metal precipitation in a zinc-contaminated, aerobic
sandy aquifer by means of biological sulfate reduction. Environ. Sci. Technol. 38(14): 4002–4011. https://doi.
org/10.1021/es030131a.
Johnson, D. B. and A. L. Santos. 2020. Biological removal of sulfurous compounds and metals from inorganic
wastewaters. pp. 215–246. In: P. Lens [ed.]. Environmental Technologies to Treat Sulfur Pollution: Principles
and Engineering (Second Edition). IWA Publishing.
Kaksonen, A. H. and J. A. Puhakka. 2007. Sulfate reduction based bioprocesses for the treatment of acid mine
drainage and the recovery of metals. Eng. Life Sci. 7(6): 541–564. https://doi.org/10.1002/elsc.200720216.
Karpiński, P. H. and J. Bałdyga. 2019. Chapter 8: precipitation processes. pp. 216–265. In: A. S. Myerson, D. Erdemir
and A. Y. Lee. [eds.]. Handbook of Industrial Crystallization (3rd Edition). Cambridge University Press. https://
www-cambridge-org.qe2a-proxy.mun.ca/core/books/handbook-of-industrial-crystallization/precipitation-pro
cesses/011386C77A45C4AAAFD4DE1B2FE0D609.
Kawaguchi, T. and A. W. Decho. 2002. A laboratory investigation of cyanobacterial extracellular polymeric secretions
(EPS) in influencing CaCO3 polymorphism. J. Cryst. Growth. 240(1): 230–235. https://doi.org/10.1016/
S0022-0248(02)00918-1.
Kiran, M. G., K. Pakshirajan and G. Das. 2017. An overview of sulfidogenic biological reactors for the simultaneous
treatment of sulfate and heavy metal rich wastewater. Chem. Eng. Sci. 158: 606–620. https://doi.org/10.1016/j.
ces.2016.11.002.
Kiskira, K., S. Papirio, E. D. van Hullebusch and G. Esposito. 2017. Fe(II)-mediated autotrophic denitrification:
a new bioprocess for iron bioprecipitation/biorecovery and simultaneous treatment of nitrate-containing
wastewaters. Int. Biodeterior. Biodegrad. 119: 631–648. https://doi.org/10.1016/j.ibiod.2016.09.020.
Kosolapov, D. B., P. Kuschk, M. B. Vainshtein, A. V. Vatsourina, A. Wießner, M. Kästner and R. A. Müller. 2004.
Microbial processes of heavy metal removal from carbon-deficient effluents in constructed wetlands. Eng.
Life Sci. 4(5): 403–411. https://doi.org/10.1002/elsc.200420048.
Kumar, N., R.-M. Couture, R. Millot, F. Battaglia-Brunet and J. Rose. 2016. Microbial sulfate reduction enhances
arsenic mobility downstream of zerovalent-iron-based permeable reactive barrier. Environ. Sci. Technol.
50(14): 7610–7617. https://doi.org/10.1021/acs.est.6b00128.
Kumar, R., M. Nongkhlaw, C. Acharya and S. R. Joshi. 2013. Bacterial community structure from the perspective
of the uranium ore deposits of domiasiat in India. Proceedings of the National Academy of Sciences, India
Section B: Biological Sciences. 83(4): 485–497. https://doi.org/10.1007/s40011-013-0164-z.
Kuppusamy, S., N. R. Maddela, M. Megharaj and K. Venkateswarlu. 2020. Total Petroleum Hydrocarbons:
Environmental Fate, Toxicity, and Remediation. Springer International Publishing. https://doi.org/10.1007/978
3-030-24035-6.
LaGrega, M. D., P. Buckingham and J. Evans. 1994. Hazardous Waste Management (Second Edition). Waveland
Press, Inc.
Levett, A., E. J. Gagen, Y. Zhao, P. M. Vasconcelos and G. Southam. 2020. Biocement stabilization of an experimental-
scale artificial slope and the reformation of iron-rich crusts. Proceedings of the National Academy of Sciences.
117(31): 18347–18354. https://doi.org/10.1073/pnas.2001740117.
Lewis, A. 2017. Precipitation of heavy metals. pp. 101–120. In: E. R. Rene, E. Sahinkaya, A. Lewis and P. N. L. Lens
[eds.]. Sustainable Heavy Metal Remediation: Volume 1: Principles and Processes. Springer International
Publishing. https://doi.org/10.1007/978-3-319-58622-9_4.
Liamleam, W. and A. P. Annachhatre. 2007. Electron donors for biological sulfate reduction. Biotechnol. Adv. 25(5):
452–463. https://doi.org/10.1016/j.biotechadv.2007.05.002.
38 Bioremediation for Sustainable Environmental Cleanup
Lookman, R., M. Verbeeck, J. Gemoets, S. Van Roy, J. Crynen and B. Lambié. 2013. In-situ zinc bioprecipitation by
organic substrate injection in a high-flow, poorly reduced aquifer. J. Contam. Hydrol. 150: 25–34. https://doi.
org/10.1016/j.jconhyd.2013.03.009.
Lovley, D. R. 1993. Dissimilatory metal reduction. Annu. Rev. Microbiol. 47(1): 263–290. https://doi.org/10.1146/
annurev.mi.47.100193.001403.
Miao, Z., M. L. Brusseau, K. C. Carroll, C. Carreón-Diazconti and B. Johnson. 2012. Sulfate reduction in groundwater:
characterization and applications for remediation. Environ. Geochem. Health. 34(4): 539–550. https://doi.
org/10.1007/s10653-011-9423-1.
Mihelcic, J. R. and J. B. Zimmerman. 2014. Environmental Engineering: Fundamentals, Sustainability, Design
(Second Edition). John Wiley & Sons, Inc.
Mitchell, J. K. and K. Soga. 2005. Chapter 8: soil deposits—their formation, structure, geotechnical properties, and
stability. pp. 195–250. In: Fundamentals of Soil Behavior (3rd Edition). John Wiley & Sons, Inc.
Mujah, D., M. Shahin and L. Cheng. 2016. State-of-the-art review of biocementation by microbially induced calcite
precipitation (MICP) for soil stabilization. Geomicrobiol. 34: 524–537. https://doi.org/10.1080/01490451.20
16.1225866.
Mwandira, W., K. Nakashima and S. Kawasaki. 2022. Chapter 13: stabilization/solidification of mining waste
via biocementation. pp. 201–209. In: D. C. W. Tsang and L. Wang [eds.]. Low Carbon Stabilization and
Solidification of Hazardous Wastes. Elsevier. https://doi.org/10.1016/B978-0-12-824004-5.00014-1.
Neculita, C. M. and G. J. Zagury. 2008. Biological treatment of highly contaminated acid mine drainage in batch
reactors: long-term treatment and reactive mixture characterization. J. Hazard Mater. 157(2): 358–366. https://
doi.org/10.1016/j.jhazmat.2008.01.002.
Nielsen, A. E. 1964. Kinetics of Precipitation. Pergamon Press.
Ossai, I. C., A. Ahmed, A. Hassan and F. S. Hamid. 2020. Remediation of soil and water contaminated with petroleum
hydrocarbon: a review. Environ. Technol. Innov. 17: 100526. https://doi.org/10.1016/j.eti.2019.100526.
Pachaiappan, R., L. Cornejo-Ponce, R. Rajendran, K. Manavalan, V. Femilaa Rajan and F. Awad. 2022. A review on
biofiltration techniques: recent advancements in the removal of volatile organic compounds and heavy metals
in the treatment of polluted water. Bioeng. 13(4): 8432–8477. https://doi.org/10.1080/21655979.2022.2050538.
Pagnanelli, F., C. C. Viggi, S. Mainelli and L. Toro. 2009. Assessment of solid reactive mixtures for the development
of biological permeable reactive barriers. J. Hazard. Mater. 170(2): 998–1005. https://doi.org/10.1016/j.
jhazmat.2009.05.081.
Reinertsen, S. A., L. F. Elliott, V. L. Cochran and G. S. Campbell. 1984. Role of available carbon and nitrogen
in determining the rate of wheat straw decomposition. Soil Biol. Biochem. 16(5): 459–464. https://doi.
org/10.1016/0038-0717(84)90052-X.
Sahinkaya, E., D. Uçar and A. H. Kaksonen. 2017. Bioprecipitation of metals and metalloids. pp. 199–231. In:
E. R. Rene, E. Sahinkaya, A. Lewis and P. N. L. Lens [eds.]. Sustainable Heavy Metal Remediation:
Volume 1: Principles and Processes. Springer International Publishing. https://doi.org/10.1007/978-3-319
58622-9_7.
Sánchez-Andrea, I., J. L. Sanz, M. F. M. Bijmans and A. J. M. Stams. 2014. Sulfate reduction at low pH to remediate
acid mine drainage. J. Hazard. Mater. 269: 98–109. https://doi.org/10.1016/j.jhazmat.2013.12.032.
Sánchez-Andrea, I., A. J. M. Stams, J. Weijma, P. Gonzalez Contreras, H. Dijkman, R. A. Rozendal and D. B.
Johnson. 2016. A case in support of implementing innovative bio-processes in the metal mining industry.
FEMS Microbiol. Lett. 363(11): 1–4. https://doi.org/10.1093/femsle/fnw106.
Sharma, H. and K. Reddy. 2004. Geoenvironmental Engineering: Site Remediation, Waste Containment and Emerging
Waste Management Technologies. John Wiley & Sons.
Shmaefsky, B. R. 2020. Principles of phytoremediation. pp. 1–26. In: B. R. Shmaefsky [ed.]. Phytoremediation:
In-situ Applications. Springer International Publishing. https://doi.org/10.1007/978-3-030-00099-8_1.
Speece, R. E. 1983. Anaerobic biotechnology for industrial wastewater treatment. Environ. Sci. Technol.
17(9): 416–427.
Stephen, J. R. and S. J. Macnaughtont. 1999. Developments in terrestrial bacterial remediation of metals. Curr. Opin.
Biotechnol. 10(3): 230–233. https://doi.org/10.1016/S0958-1669(99)80040-8.
UNDP. 2022. Sustainable development goals: United Nations Development Programme. UNDP. https://www.undp.
org/sustainable-development-goals.
Vanbroekhoven, K., S. Van Roy, L. Diels, J. Gemoets, P. Verkaeren, L. Zeuwts, K. Feyaerts and F. van den Broeck.
2008. Sustainable approach for the immobilization of metals in the saturated zone: in situ bioprecipitation.
Hydrometall. 94(1–4): 110–115. https://doi.org/10.1016/j.hydromet.2008.05.048.
Waybrant, K. R., C. J. Ptacek and D. W. Blowes. 2002. Treatment of mine drainage using permeable reactive barriers:
column experiments. Environ. Sci. Technol. 36(6): 1349–1356. https://doi.org/10.1021/es010751g.
Bioprecipitation as a Bioremediation Strategy for Environmental Cleanup 39
Webb, J. S., S. McGinness and H. M. Lappin-Scott. 1998. Metal removal by sulphate-reducing bacteria from natural and
constructed wetlands. J. Appl. Microbiol. 84(2): 240–248. https://doi.org/10.1046/j.1365-2672.1998.00337.x.
Wilcox, S. M., C. N. Mulligan and C. M. Neculita. 2022. Bioprecipitation as a remediation technique for metal(loid)
contamination from mining activities. In: A. Malik, M. K. Kidwai and V. K. Garg [eds.]. Bioremediation of
Toxic Metal(loid)s. CRC Press, Taylor and Francis group.
Willis, G. and E. R. Donati. 2017. Heavy metal bioprecipitation: use of sulfate-reducing microorganisms.
pp. 114–130. In: E. R. Donati [ed.]. Heavy Metals in the Environment: Microorganisms and Bioremediation.
CRC Press. https://doi.org/10.1201/b22013.
Xiangliang, P. 2009. Micrologically induced carbonate precipitation as a promising way to in situ immobilize heavy
metals in groundwater and sediment. Res. J. Chem. Environ. 13(4): 3–4.
Yong, R. N., C. N. Mulligan and M. Fukue. 2014. Sustainable Practices in Geoenvironmental Engineering (2nd ed.).
CRC Press. https://doi.org/10.1201/b17443.
Zagury, G. J., V. I. Kulnieks and C. M. Neculita. 2006. Characterization and reactivity assessment of organic substrates
for sulphate-reducing bacteria in acid mine drainage treatment. Chemosphere. 64(6): 944–954. https://doi.
org/10.1016/j.chemosphere.2006.01.001.
Chapter 3
Bio Adsorption
An Eco-friendly Alternative for
Industrial Effluents Treatment
Andrea Saralegui,1,* M. Natalia Piol,1
Victoria Willson,1 Néstor Caracciolo,2 Silvia Ramos 3
and Susana Boeykens1
3.1 Introduction
Industrial effluents are complex systems and therefore their treatment requires an integral approach.
The composition and characteristics of each type of effluent, whether solid or liquid, determine the
treatment to be required and its final disposal. In order to find a feasible solution for each specific
case, it is necessary to recognize the system and look for alternatives that reduce costs.
Both waste and pollutants can be considered as misplaced or unexploited resources. That is to
say, the lack of a process’s holistic view means that an incorrect disposal of a material turns it into a
waste or even a pollutant. The reuse of this material not only reduces the degree of contamination,
but also reduces the operational costs of its disposal and/or treatment. Furthermore, it turns it into a
valuable resource that could even have some added value in other activities. Thus, what was earlier
called “waste” could be seen as a new raw material in a recycling and reuse process, becoming a loop
within the circular economy (Saralegui et al. 2022). The study of these new sources of raw materials
is a way forward to improve the sustainability of water and effluent treatment technologies.
Environmental contamination with metals can negatively affect ecosystems and the exposed
population health; this contamination is mainly due to technological changes and the increasing
use of metal-containing materials in industry. This has led to concern and has led to numerous
investigations on treatment technologies for water, effluents and contaminated soils (Branzini
and Zubillaga 2012, Cartaya et al. 2011, EPA 2000, Volke Sepúlveda and Velasco Trejo 2002).
1
Universidad de Buenos Aires, Facultad de Ingeniería, Instituto de Química Aplicada a la Ingeniería (IQAI), Laboratorio de
Química de Sistemas Heterogéneos (LaQuíSiHe). Av. Paseo Colón 850, C1063ACV, Buenos Aires, Argentina.
2
Universidad de Buenos Aires, Facultad de Ingeniería, Instituto de Química Aplicada a la Ingeniería (IQAI), Laboratorio de
Química Ambiental (LaQuíAmb). Av. Paseo Colón 850, C1063ACV, Buenos Aires, Argentina.
3
Universidad de Buenos Aires, Facultad de Ingeniería, Grupo Modelos Aplicados a Gestión Industrial, Dpto. de Gestión-
Cátedras de Modelos y Optimización I y III. Av. Las Heras 2214, C1126, Buenos Aires, Argentina.
* Corresponding author: [email protected]
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 41
Aquatic systems are large bodies that receive many waste streams, mainly resulting from human
activities (Acumar 2019, Liu et al. 2008). The diversity of industries that may discharge their effluents
into a water body influences its degree and type of pollution. The assessment of the interferences that
may exist between various types of pollutants present simultaneously in a water body should keep
being studied as they may produce different effects on the ecosystem (Renaud et al. 2021). Prevention
is the way forward. Avoiding the discharge of effluents with high pollutant content is necessary and
finding systems that could selectively remove or decrease a specific pollutant concentration in the
presence of other substances is very promising regarding their recovery and reuse.
Due to their high persistence in the environment, metals have the potential to get bioaccumulated
in organisms of lower trophic levels and reach their biomagnification, with a direct implication
on the ecosystem and on the health of populations not directly exposed to contaminated water
(Di Giulio and Newman 2008). For these reasons, the Argentinian legislation considers them as
hazardous substances and regulates their content both in effluent discharges and in drinking and
irrigation waters (PLNRA 1992). However, this legislation as well as other water quality regulations
and guidelines in the world do not consider the effect of the simultaneous presence of different
metals (Renaud et al. 2021).
The search for low-cost processes for water treatment is fundamental to its applicability; waste
reuse is a possible path to follow. When thinking about the low cost of the process, this implies that
the waste should be sourced close to where it will be used in order to reduce transport expenses
as well as the carbon footprint of the process. It also implies making just a basic conditioning to
the waste, that is, treating the material as little as possible. Finally, the process must be adapted
to the specific effluent conditions, so as not to require adjustments involving the use of large
quantities of chemical reagents or energy. There are several established conventional processes for
pollutants treatment and recovery from wastewater (Mihelcic and Zimmerman 2012). However, in
recent years, studies on biosorption technologies development have increased due to their potential
economic convenience. Numerous works using lignocellulosic materials indicate a high capacity
to concentrate water pollutants in their structures (Boeykens et al. 2018, Boeykens et al. 2019, Piol
et al. 2021, Saralegui et al. 2021, Saralegui et al. 2022). For this work, wastes with high availability
in Argentina were selected: peanut shells (Arachis hypogaea), sugar cane bagasse (Saccharum
officinarum), avocado stones (Persea americana), pecan nut shells (Carya illinoinensis), wheat
bran (Triticum aestivum), banana shells (Mussa paradisiaca) and different parts of the moringa
plant (Moringa oleifera) for the study of metal removal from water.
Selected wastes:
• Peanut shells: Argentina is the sixth-largest producer of peanuts in the world, accounting
for 3% of global production, with 42.6 million tons produced during the 2016–2017 growing
season (Pellegrino 2019). Peanut shells, the waste from the industrialization of the nuts, are
often a drawback as they represent between a quarter and a fifth of the harvest and constitute a
polluting waste which is usually incinerated in the open air, generating large amounts of smoke
(BCC 2014).
• Sugar cane bagasse: another waste product is sugar cane bagasse, which is the fibrous
material that remains after the sugar cane is crushed. Typically, for every 10 tons of crushed
cane, 3 to 4 tons of wet bagasse (40–50% moisture) remains as residue. Increasingly, it is used
as fuel for steam and electricity generation in sugar mills, but also as a source of organic matter
when returned to the soil. According to statistics, in the last 10 yr’ harvest, 18 to 20 million
tons of sugar cane were milled in the country, producing a large volume of bagasse for disposal
(Preciado Patiño 2015, Rios et al. 2017).
• Avocado stone: the avocado stone represents about 18% of the fruit weight. In Argentina
it is currently a growing crop as different soils and climates are favorable for its growth and
development (Carrere 2010).
42 Bioremediation for Sustainable Environmental Cleanup
• Pecan nuts: these agronomic conditions also promote the production of high quality, counter-
seasonal pecan nuts. Commercial pecan cultivation in Argentina has grown exponentially
in recent years, which in the medium term will make Argentina one of the world’s top three
producers of this nut and the world’s leading exporter of high value-added pecan-based products
(Frusso 2013).
• Wheat: On the other hand, wheat is one of the most widely grown cereals in the world. In
Argentina, around 15.5 million tons of wheat are produced annually. Seventy-five per cent is
exported and approximately 4.6 million tons are milled to produce 3.4 million tons of flour.
As by-products of this process, approximately 500,000 tons of bran and germ are extracted
annually and used almost entirely for animal feed (Apro et al. 2004).
• Banana: Banana cultivation is the most prominent of Argentina’s tropical crops, with an annual
production ranging from 180,000 to 205,000 tons, with bananas being the most consumed fruit
in Argentina with an annual per capita average of 12 kg (Molina et al. 2015).
• Moringa oleifera: Moringa oleifera, a shrubby plant native to India, is little known in
South America, although it is now abundant throughout the tropics. It is a hardy species that
requires little horticultural attention and grows rapidly, recognized for its nutritional and
medicinal characteristics, as well as being used in some water purification processes (Folkard
and Sutherland 1996). Different parts of the plant are usable raw materials, the leaves, the
undeveloped pods and seeds. Obtaining products and by-products generates around 80% by
crop weight of waste, the husks during dehulling the seed or the green stalks are discarded
during the obtaining of the leaf for infusions use. Currently, in Argentina there is a growing
wave of Moringa cultivation with a large generation of new products that in the short term will
result in a large amount of waste to be disposed of.
The waste from the agri-food industry proposed in this work will be used as biosorbents to
study the retention of metal ions from aqueous effluents. In the case of liquid effluents, the aim
is to reach the polluting discharge levels required by local legislation, considering industries such
as tanning or electroplating. In the case of solid effluents, the goal is to achieve eco-friendly final
disposal.
The design of the Reactor App software were intended to simplify the calculations and to
give as a result the volume of the continuous reactor to be used in a first approximation from data
obtained with few tests carried out on batch reactors. In this way, a first scale-up can be carried out
in a simplified way. As a second objective, Reactor App aims to facilitate the processing of the data
derived from the continuous tests.
Figure 3.1. Peanut shells before and after washing, drying and sieving.
Figure 3.2. Peanut shell dosage curves. a. As a function of adsorbent mass. b. As a function of initial nickel solution
concentration.
The compromise relation selected was 0.3 g of adsorbent and 20 mg Ni2+ L–1 as the initial
concentration. For this ratio the percentage of removal obtained was close to 80% and the adsorption
capacity close to the maximum. This compromise relation also allows measurable results to be
obtained within the concentration range suitable for the measuring instrument.
44 Bioremediation for Sustainable Environmental Cleanup
Figure 3.3. Kinetic curves obtained from the adsorption of nickel on peanut shells.
Once the ratio between the working concentration and the mass of the adsorbent to be used
was established, the kinetic tests begin with the objective of obtaining the kinetic constant and the
reaction order by applying different models.
As an example of the kinetic studies, the kinetic curves for 10, 20 and 30 mg Ni L–1, using 0.3 g
of adsorbent, are presented in Figure 3.3. The equilibrium time for all systems was 30 min.
By processing these studies one can determine the order of the reaction and the kinetic constant.
For this, it is necessary to use pseudo-first order and pseudo-second order models. The mathematical
expression for the pseudo-first order kinetics model (Lagergren 1898) is widely used for adsorption
studies of liquids. The linear form is Eq. 3.3:
where qt (mmol g–1) is the amount of adsorbed solute at time t, and k (min–1) is the constant velocity.
The pseudo-second order model was developed by Ho and McKay (1999). This model assumes
that adsorption follows a second-rate kinetic mechanism. It indicates that one sorbate molecule is
adsorbed on two active sites on the sorbent surface. The process can be expressed in a linear form
by the Eq. 3.4:
t 1 t
= + t Eq. 3.4
qt k2 qe2 qe
where, k2 (mmol g–1 min–1) is the velocity constant of pseudo-second order. The term (k 2 qe2)
represents the initial adsorption rate.
In the case of peanut shells, the best fit was obtained for the pseudo-second order model
with regression coefficients higher than 0.998, while for the pseudo-first order fit the regression
coefficient in all cases was less than 0.92. In general, the reaction order for lignocellulosic materials
for metal removal responds to reaction of order 2. Several authors have demonstrated that the metal
ions adsorption process in lignocellulosic materials occurs by complexation, it occurs where the
carboxylic or carboxylate (R-COOH/R-COO-) and phenolic or phenolate (R’-OH / R’-O-) functional
groups simultaneously chelate the metal ions (Mn+), forming bidentate complexes R-COO-Me-
O-R’ (Neris et al. 2019, Guo et al. 2008, Piol et al. 2021, Boeykens et al. 2018).
Regarding the kinetic constant obtained, values between 600 and 15 g mmol–1 were obtained.
The cyclic of these parameters (k and reaction order) is necessary to enter the Reactor App software.
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 45
Figure 3.4. Experimental data and non-linear fit of the used models to the copper on Azolla adsorption isotherms.
Table 3.2. Obtained parameters from the adsorption isotherms fit to both models.
Langmuir Freundlich
Adsorbent-adsorbate qmáx KL R2 KF n R2
Azolla-Cu 0.457 14.19 0.9837 0.399 4.70 0.8895
amount of metal that is possible to retain. In this way, a successful design of a continuous adsorption
reactor requires being able to predict the characteristics of the breakthrough curve (Acheampong
et al. 2013). The development of mathematical models describing the dynamic behavior of adsorption
in fixed beds is difficult, because the adsorbate concentration, like the feed, moves through the
column and a steady state cannot be assumed.
During the application of a model that approximates the experimental parameters, several
factors should be considered, among them the adsorbate transport in fixed beds, which is described
by the solid-liquid material balance equations. This description must account for the mass balance
of the adsorbed solute, which in turn depends on the mechanism responsible for adsorption and
may be controlled by mass transfer from within the solution to the adsorbent surface or by the
chemical reaction between the adsorbent particles and the metal (Calero et al. 2009). In this sense,
determining the parameters of models of varying complexity can be difficult, considering the large
number of resources required and the complicated solution methods (Borba et al. 2008). This is
why several simplified mathematical models have been developed, which provide parameters that
qualitatively describe the effects during adsorption in the continuous reactor (Chu et al. 2007).
Among them, the most widely applied models are Thomas (1944), Bohart-Adams (1920) and Yoon
and Nelson (1984), which will be mentioned in the present investigation. Each model presents
different mathematical equations so that from each, different parameters are obtained that provide
different information about the processes under study. The assumptions and parameters of each
model are shown in Table 3.3.
For the purpose of checking the applicability of the described models, the parameters provided
by the different models (Xcal) should be compared with the experimentally obtained parameters
Table 3.3. Comparison between the used models for continuous reactors.
Parameters Graph C/Co Vs Vef Graph C/Co Vs Vef Graph C/Co Vs Vef KYN
KTH is Thomas rate constant KAB is the mass transfer stands for the Yoon-Nelson
[mL min–1 g–1], coefficient [cm3 mmol–1 min–1], rate constant [min–1],
q0 is the maximum solute N0 represents the maximum τ [min] is the time
concentration in the solid phase adsorption capacity [mmol cm–3], required to retain 50% of
[mmol g–1], U is the linear liquid velocity the C0.
Vef is the effluent volume [L], [cm min–1],
F is the volumetric flow rate [mL Z is the bed height in the reactor
min–1], [cm].
W is the amount of sorbent inside the
reactor [g],
C is the outlet concentration
[mmol L–1].
Assumptions ● Adsorption behavior follows the ● Adsorption rate is proportional ● Adsorption
Langmuir isotherm to the residual capacity of rate decreases
● Intraparticle diffusion and the solid and the adsorbate proportionally with the
resistance to external mass concentration number of molecules
transfer are negligible. ● Intraparticle diffusion and adsorbed.
resistance to external mass
transfer are negligible.
48 Bioremediation for Sustainable Environmental Cleanup
(Xexp) by means of goodness-of-fit analysis, which is evaluated by the regression coefficient R2 and
the average relative error (ARE) (Dhanasekaran et al. 2017) defined by the following expression:
(X i,calc − X i,exp )
ARE = 100
— n
Eq. 3.5
n ∑i =1 X i,exp
The proximity of the value of R2 to one (Tan and Hameed 2017) and ARE values lower than
35% (Tsai et al. 2016) indicate a good fit of the model to the experimental data.
To exemplify the application of these models, the work system was compounded by Azolla
biomass with a contaminant solution of copper, for which a continuous reactor was assembled,
whose dimensions were 15 cm in height, 15.9 cm3 in volume and an internal diameter of 1.161 cm,
which was completed with 4.3 g of biomass with a particle size between 1.18–0.5 mm and then a
Cu(II) solution of 50 mM concentration was passed upstream at a fixed flow rate of 0.5 mL min–1
and the Cu(II) concentration at the reactor outlet was measured at different volumes by ultraviolet-
visible spectrophotometry. From the breakthrough curves obtained (C/C0 versus V graph) the
nonlinear fitting models mentioned above were applied and the corresponding parameters listed in
Table 3.4 were obtained.
Table 3.4. Obtained parameters with the Thomas–Bohart-Adams and Yoon-Nelson models for the adsorption of copper onto
Azolla biomass in fixed bed reactor.
Azolla-Cu
KTH 2.39
q 0 calc 0.6222
Thomas
q 0 exp 0.679
Model
R2 0.996
ARE 8.39
KAB 1.637
No calc 0.1951
Bohart-Adams Model No exp 8.68E-2
R2 0.9770
ARE 125
Kyn 0.1005
τcalc 136.3
Yoon-Nelson Model τexp 135.0
R2 0.996
ARE 0.96
From the data obtained for the modeling of the experimental breakthrough curve it can be
observed that both the Thomas and Yoon-Nelson models are the most appropriate to describe the
behavior of the experimental data, since both have R2 greater than 0.99 and an ARE less than 35%,
which would indicate that the value of the parameters calculated by the model and the experimental
one are very similar. These models adequately fit adsorption processes where external and
internal diffusions are not the limiting step (Aksu and Gönen 2004). Both models have analogous
mathematical equations, so they were expected to predict similar fits (Chu 2020), however, from
each of them, different information about the system under study can be obtained. The Thomas
model allows estimating the maximum amount of adsorbate retained in the solid phase, from the
calculated q0 which resulted to be 0.622 mmol g–1. On the other hand, the Yoon-Nelson model
allows knowing the value of the time required for the concentration in the effluent to be equal to
50% of the input, the latter, determined by the parameter τ, was 136 min.
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 49
The successful design of a fixed-bed column adsorption process requires the prediction of
the effluent breakthrough curve (Chen et al. 2012), so the parameters obtained together with the
adsorbent saturation times are very useful and represent a starting point for future studies. The
description of these breakthrough curves was shown in the work presented by Saralegui et al. (2022).
The optimization of a continuous reactor, varying parameters such as the amount of each adsorbent,
flow rate, column size, among others are the subject of further studies.
Calculation of curves
Search adsorbent Reactor volume areas Breakthrough curve
models
Data on adsorbents, adsorbates and pre-loaded adsorbent-adsorbate systems are needed to start
the task. The software works through cards for each adsorbate (Figure 3.7) or adsorbent (Figure 3.8)
that are pre-filled. These cards include the physicochemical characteristics of both the adsorbate and
the adsorbent. These data can be obtained from literature (in the case of adsorbates, legislation or
guide levels are included) or empirically and are used for the implementation of the models included
in the software.
Figure 3.8 shows the cards corresponding to the systems or adsorbate-adsorbent pairs studied.
In these cards, the data obtained in batch tests such as qmax and Langmuir constant, equilibrium time,
reaction order and kinetic constant, among others are loaded.
Figure 3.9 shows the screen where the information to calculate the first approximation of the
volume of a continuous reactor must be loaded.
Reactor App also allows uploading experimental data from continuous reactors (by uploading a
CSV or XLSX file), both from fluid dynamic tests in which inert material is used as the reactor filler
and additionally tests with the reactor filled with the adsorbent under study. By difference of the
areas under the breakthrough curve (Saralegui et al. 2022), in both cases the amount of contaminant
removed by the reactor can be calculated (Figure 3.10).
With the experimental data of the breakthrough curves, Reactor App is able to perform the
estimation of different models used as approximations and return the characteristic fitting parameters
of each one together with the regression coefficient R2. The models that are possible to apply are
Thomas, Bohart-Adams and Yoo-Nelson, explained above, and an example from the Reactor App
return page is shown in Figure 3.11.
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 51
(a)
(b)
Figure 3.7. Cards from the Reactor App software corresponding to the studied (a). Adsorbates (Text in Spanish indicating
their different parameters such as ionic charge, hydrodynamic radio and discharge limit) and (b). Adsorbents (Text in Spanish
indicating their particle size, BET area and pH).
52 Bioremediation for Sustainable Environmental Cleanup
Figure 3.8. Cards of the Reactor App software corresponding to the systems (adsorbate/adsorbent pairs) studied (Text in
Spanish indicating some parameters such as equilibrium time, qmax, temperature and initial pH).
Figure 3.9. Data to be entered in Reactor App for the calculation of the first continuous reactors volume approximation.
(Text in Spanish indicating the different variables for the selected system: adsorbate, adsorbent, equilibrium time, qmax,
temperature, initial pH, kinetic constant and reaction order).
3.7 Conclusions
In conclusion, the choice of the low-cost adsorbent to be used should be based on the availability
of the material in the same place where the effluent treatment is needed. In order to simplify the
studies, first the dosage curve should be made and the working conditions should be established.
Once the relation between the contaminant concentration and the quantity of adsorbent are fixed,
the kinetic equation at a given temperature should be estimated. Then, the study of the equilibrium
isotherm and the estimation of the maximum adsorption capacity could be accomplished. Finally,
with this data the continuous reactor volume can be estimated using Reactor App and then a first
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 53
Figure 3.10. Graph obtained in Reactor App by entering the data of the breakthrough curves of the fluid dynamic test (pink
dots) and the adsorption of Lead (II) on dolomite (light blue dots). The red area represents the amount of adsorbed lead, the
text in Spanish additionally indicates the base line area, the difference between these both areas and the adsorbed amount in
mmol.
Figure 3.11. Example of the Reactor App return screen for the application of the Thomas model for the data obtained for
copper (II) adsorption on Azolla pinnata in the continuous reactor.
approximation for the scale-up can be made. Once the continuous reactor is built, the results can be
analyzed with Reactor App and the models can be used to parameterize the new reactor. In this way
it is possible to consider a new scale jump.
In all the cases studied in our laboratory, the obtained results indicate that the use of
lignocellulosic materials for the removal of ionic pollutants from wastewater is feasible, taking into
account that the use of these materials is a contribution to the circular economy.
We are working on the mixing of contaminants and multiple adsorptions on these adsorbents,
as well as on the safe final disposal of these solids when exhausted.
54 Bioremediation for Sustainable Environmental Cleanup
Acknowledgements
The authors acknowledge the financial support from Universidad de Buenos Aires (UBACyT
N°20020190100323BA, N°20020190200302BA, PDE 032/2020) and would also like to thank the
computer engineering students Matías Reimondo, Santiago Pinto and Lucas Lavandeira for their
collaboration in the development of the application used in this work.
References
Acheampong, M., K. Pakshirajan, A. Annachhatre and P. Lens. 2013. Removal of Cu(II) by biosorption onto coconut
shell in fixed-bed column systems. J. Ind. Eng. Chem. 19: 841–848.
ACUMAR. 2019. Records of establishments in the Matanza Riachuelo Basin 2019(in Spanish). Retrieved from http://
datos.acumar.gob.ar/dataset/registro-de-establecimientos-de-la-cuenca-matanza-riachuelo-2019/archivo/
dca9d704-8e4c-4bbb-bed1-2a7460788eb0 consulta 3/08/2019.
Aksu, Z. and F. Gönen. 2004. Biosorption of phenol by immobilized activated sludge in a continuous packed bed:
prediction of breakthrough curves. Process Biochem. 39: 599–613.
Apro, N. J., C. J. Cuadrado and P. A. Secreto. 2004. Stabilization of wheat bran and germ, through the extrusion
process as an input for the food industry (in Spanish). Development and Innovation Day. INTI-Cereals and
Oilseeds.
BCC, Bolsa de comercio de Córdoba. 2014. Chapter 15: Peanut production chain (in Spanish). Retrieved from https://
bolsacba.com.ar/buscador/?p=1354.
Boeykens, S. P., M. N. Piol, L. Samudio Legal, A. B. Saralegui and C. Vázquez. 2017. Eutrophication decrease:
phosphate adsorption processes in presence of nitrates. J. Environ. Manage. 203: 888–895.
Boeykens, S. P., A. B. Saralegui, N. Caracciolo and M. N. Piol. 2018. Agroindustrial waste for lead and chromium
biosorption J. Sustain. Dev. Energy Water Environ. Syst. 6: 341–350.
Boeykens, S. P., N. Redondo, R. A. Obeso, N. Caracciolo and C. Vázquez. 2019. Chromium and lead adsorption by
avocado seed biomass study through the use of Total Reflection X-Ray Fluorescence analysis. Appl. Radiat.
Isot. 153: 108809.
Bohart, G. and E. Adams. 1920. Some aspects of the behavior of charcoal with respect to chlorine. J. Am. Chem. Soc.
42: 523–544.
Borba, C., E. Da Silva, M. Fagundes-Klen, A. Kroumov and R. Guirardello. 2008. Prediction of the copper (II) ions
dynamic removal from a medium by using mathematical models with analytical solution. J. Hazard. Mater.
152: 366–372.
Branzini, A. and M. Zubillaga. 2012. Remediation and monitoring of soils contaminated with heavy metals: Organic
and inorganic amendments to remediate soils contaminated with metals (in Spanish). Editorial Académica
Española.
Calero, M., F. Hernáinz, G. Blázquez, G. Tenorio and M. Martín-Lara. 2009. Study of Cr (III) biosorption in a fixed-
bed column. J. Hazard. Mater. 171: 886–893.
Carrere, R. 2010. The avocado: a fruit tree for the family garden (in Spanish). Retrieved from https://docplayer.es/
amp/14007991-La-palta-un-frutal-para-la-huerta-familiar.html.
Cartaya, O. E, I. Reynaldo, C. Peniche and M. L. Garrido. 2011. Use of natural polymers as an alternative for the
remediation of soils contaminated by heavy metals (in Spanish). Rev. Int. de Contam. Ambient. 27: 41–46.
Chen, S., Q. Yue, B. Gao, Q. Li, X. Xu and K. Fu. 2012. Adsorption of hexavalent chromium from aqueous solution by
modified corn stalk: a fixed-bed column study. Bioresource Technol. 113: 114–120. https://doi.org/10.1016/j.
biortech.2011.11.110.
Chu, K. H. and M. A. Hashim. 2007.Copper biosorption on immobilized seaweed biomass: Column breakthrough
Characteristics. J. Environ. Sci. 19: 928–932.
Chu, K. H. 2020. Breakthrough curve analysis by simplistic models of fixed bed adsorption: in defense of the century-
old Bohart-Adams model. Chem. Eng. J. 380: 122513. https://doi.org/10.1016/j.cej.2019.122513.
Dhanasekaran P., P. Satya and K. I. Gnanasekar. 2017. Fixed bed adsorption of fluoride by Artocarpus hirsutus based
adsorbent. J. Fluor. Chem. 195: 37–46.
Di Giulio, R. T. and M. C. Newman. 2008. Ecotoxicology. In Casarett and Doull’s. Toxicology. The basic science of
poisons, ed. C. D. Klaasen. Kansas: McGraw-Hill.
EPA, Environmental Protection Agency. 2000. In situ treatment of soil and groundwater contaminated with
chromium-technical resource. 625/R-00/004. Retrieved from https://cfpub.epa.gov/si/si_public_record_
report.cfm?Lab=NRMRL&dirEntryId=64150.
Folkard, G. and J. Sutherland. 1996. Moringa oleifera: A tree with enormous potential (in Spanish) Original in JAMA.
1931, 8(3): 211. Traducido por Ariadne Jiménez U.C.R., Turrialba, Costa Rica. Agroforestry Today, 8(3).
Bio Adsorption: An Eco-friendly Alternative for Industrial Effluents Treatment 55
Foo, K. Y. and B. H. Hameed. 2010. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J.
156: 2–10.
Freundlich, H. M. F. 1906. Über die Adsorption in Lösungen. Zeitschriftfür Physikalische Chemie (Leipzig) 57 A:
385–470.
Frusso, E. A. 2013. Influence of nitrogen, phosphorus and zinc on the chemical composition and yield of pecan nuts
and their relationship with the variability of nutrients in the leaf (in Spanish) thesis for master’s degree in Plant
Production. Universidad de Buenos Aires.
Guo, X., S. Zhang and X. Shan. 2008. Adsorption of metal ions on lignin. J. Hazard. Mater. 151: 134–142. doi:https://
doi.org/10.1016/j.jhazmat.2007.05.065.
Ho, Y. S. and G. McKay. 1999. Pseudo-second order model for sorption processes. Process. Biochem. 34: 451–465.
Lagergren, S. 1898. About the theory of so-called adsorption of soluble substances (in German: Zurtheorie der
sogenannten adsorption gelosterstoffe), Supplement of the Royal Swedish Science Academiens Documents
(in Swedish: Bihang Till Konglinga Svenska Vetenskaps Academiens Handlingar 24: 1–39.
Langmuir, I. 1918. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 40 (9):
1361–1403.
Liu, J., R. A. Goyer and M. P. Waalkes. 2008. Toxic effects of metals. In Casarett and Doull’s. Toxicology. The Basic
Science of Poisons., ed. C. D. Klaasen. New York, United State: McGraw - Hill.
Mihelcic, J. and J. Zimmerman. 2012. Environmental Engineering: Fundamentals, Sustainability, Design (in Spanish).
Alfaomega. Mexico.
Molina, N. A., F. Scribano, G. Tenaglia and D. Rodríguez. 2015. Cost of banana production in Formosa (in Spanish).
EEA Bella Vista. Serie Técnica, 50.
Neris, J. B., F. H. M. Luzardo, P. F. Santos, O. N. de Almeida and F. G. Velasco. 2019. Evaluation of single and
tri-element adsorption of Pb(II), Ni(II) and Zn(II) ions in aqueous solution on modified water hyacinth
(Eichhornia crassipes) fibers. J. Environ. Chem. Eng. 7(1): 102885. doi:https://doi.org/10.1016/j.
jece.2019.102885.
Pellegrino, M. 2019. Peanut Chain - summary. Argentina: Presidency of the Nation (in Spanish) Retrieved from https://
alimentosargentinos.magyp.gob.ar/HomeAlimentos/Cadenas%20de%20Valor%20de%20Alimentos%20
y%20Bebidas/informes/Resumen_Cadena_2018_Mani_manies_crudos_y_Preparaciones_de_mani.pdf.
Piol, M. N., C. Dickerman, M. P. Ardanza, A. B. Saralegui and S. P. Boeykens. 2021. Simultaneous removal of
chromate and phosphate using different operational combinations for their adsorption on dolomite and banana
peel. J. Environ. Manage. 288: 112463.
PLNRA. Poder Legislativo Nacional. República Argentina. 1992. National Hazardous Waste Law No. 24.051 - Decree
No. 831/93 - Resolution No. 242/93 Standards for Discharges from Industrial or Special Establishments
reached by Decree 674/89 (in Spanish) C.F.R. 1992. http://servicios.infoleg.gob.ar/infolegInternet/
anexos/10000-14999/12830/norma.htm.
Preciado Patiño, J. 2015. The Sugarcane Chain and its agro-industrial value (in Spanish). Council of Argentine Agro,
Agrifood and Agroindustry Professionals, http://www.cpia.org.ar/agropost/201509/nota1.html.
Renaud, M., P. M. da Silva, T. Natal-da-Luz, S. D. Siciliano and J. P. Sousa. 2021. Community effect concentrations
as a new concept to easily incorporate community data in environmental effect assessment of complex metal
mixtures. J. Hazard. Mater. 411: 125088. doi:https://doi.org/10.1016/j.jhazmat.2021.125088.
Ríos, L., G. Pérez and A. Felipe. 2017. The Argentine sugar market and Economic analysis of the sugar harvest in
Tucumán Campaign 2016–2017 (in Spanish). INTA.
Salamatinia, B., A. Kamaruddin and A. Abdullah. 2008. Modeling of the continuous copper and zinc removal by
sorption onto sodium hydroxide-modified oil palm frond in a fixed-bed column. J. Chem. Eng. 145: 259–266.
Saralegui, A. B., V. Willson, N. Caracciolo, M. N. Piol and S. P. Boeykens. 2021. Macrophyte biomass productivity
for heavy metal adsorption. J. Environ. Manage. 289: 112398.
Saralegui, A. B., M. N. Piol, V. Willson, N. Caracciolo and S. P. Boeykens. 2022. Lignocellulosic waste as adsorbent for
water pollutants. A step towards sustainability and circular economy. pp. 168–182. In: A. Malik, M. K. Kidwai
and V. K. Garg [eds.]. Bioremediation of Toxic Metal(loid)s : CRC press, Taylor and Francis group, Boca Raton.
Tan, K. L. and B. H. Hameed. 2017. Insight into the adsorption kinetics models for the removal of contaminants from
aqueous solutions. J. Taiwan Inst. Chem. Eng. 74: 25–48.
Thomas, H. C. 1944. Heterogeneous ion exchange in a flowing system. J. Am. Chem. Soc. 66: 1664–1666.
Tsai, W., M. De Luna, H. Bermillo-Arriesgado, C. Futalan, J. Colades and Meng-WeiWan. 2016. Competitive fixed-
bed adsorption of Pb(II), Cu(II), and Ni(II) from Aqueous Solution Using Chitosan-Coated Bentonite. Int. J.
Polym. Sci. http://dx.doi.org/10.1155/2016/1608939.
Volke Sepúlveda, T. and J. Velasco Trejo. 2002. Tecnologías de remediación para suelos contaminados. Mexico:
Instituto Nacional de Ecología (INE-SEMARNAT).
Yoon, Y. H. and J. H. Nelson. 1984. Application of gas adsorption kinetics I. A theoretical model for respirator
cartridge service life. Am. Ind. Hyg. Assoc. J. 45: 509–516.
Chapter 4
Bioaccumulation and
Biosorption
The Prospects and Future Applications
P.F. Steffi1,*1 and P.F. Mishel 2
4.1 Introduction
The rapid development of industry and modern technologies has shown the release of variable
hazardous compounds into the surroundings, including heavy metals (Qiu et al. 2021). Heavy
metals are utilized in the mining, metallurgical, electronics and electroplating sectors. Many modern
operations produce trash that contains heavy metals which are harmful to both lower and higher
living beings (Devanesan and AlSalhi 2021). There are a number of ways for removing heavy metal
ions now in use. Precipitation, ion exchange, membrane processes, evaporation and filtration are
the most significant (Danouche et al. 2021). The use of these approaches is frequently coupled
with technological challenges, such as waste management (Pushkar et al. 2021). Biotechnological
strategies that utilize biological materials such as bacteria and plants could be a viable alternative to
the chemical and physical systems currently in use (Reddy et al. 2021).
Metals can be bound by biological material through biosorption and bioaccumulation processes
(Abdul Jaffar et al. 2015). During the biosorption process, metal ions are adsorbed on the surface of
a sorbent. Biosorption is a metabolically passive energy-producing process that uses dead biomass
as a fuel source. Biosorption is a process that has a few distinct characteristics (Hansda et al. 2016).
It has great effectiveness in sequestering dissolved metals from very dilute complicated solutions. As
a result, biosorption is an excellent choice for treating high-volume, low-concentration complicated
wastewaters (Joshi et al. 2011).
On the other hand, bioaccumulation can only occur in living beings due to the transit of
pollutants into the cell and the accumulation of metals within the cell (Kumar et al. 2021). The initial
phase in bioaccumulation is biosorption. Bioaccumulative chemicals are those that accumulate in
living beings to the point where their concentrations in bodily tissues continue to rise (Kumar et al.
2007). Bioaccumulation is also known as bioconcentration in fish and other aquatic creatures. The
1
PG and Research Department of Microbiology, Cauvery College for Women (Autonomous), Trichy.
2
Department of Botany, Bharathidasan University, Trichy.
* Corresponding author: [email protected]
Bioaccumulation and Biosorption: The Prospects and Future Applications 57
general objective of this chapter is to look at the biosorption and bioaccumulation potentials for
heavy metal removal from contaminated environments (Singh and Kumar 2020). The most utilized
heavy metal removal technologies from water and wastewater are discussed.
impact on climate change, causes economic losses (mainly in agriculture and forestry) and offers a
significant health risk to humans (Cachada et al. 2014).
metal removal requires metabolic activity, bioaccumulation occurs only when microbial cells are
alive. Biosorption is a technique that involves the utilization of dead biomass. Materials utilized in
biological processes to remove metals include wasted biomass and organisms found naturally in the
environment. This lowers the cost of biological approaches, making them more accessible to a wider
range of applications (Garg et al. 2012).
4.4 Biosorption
The removal of heavy metals using live microbes is complicated. Heavy metals have a harmful
influence on living organisms’ cells, which is the main issue. Advanced research, on the other hand,
revealed that dead cells could attach metal ions through a variety of physicochemical methods
(Gerber et al. 2018). Biosorption and biosorbents of various origins have received a lot of interest as
a result of this new finding. Several studies in this field have contributed to the spread of biosorption
knowledge (Hansda et al. 2016). As a result, the number of new potential applications has increased.
This process is known to be influenced not only by the kind and chemical composition of the
biomass but also by external physicochemical factors. The processes behind the biosorption process
have also been identified and elucidated. Many research centres across the world are working to
better understand biosorption (Hlihor et al. 2015). Furthermore, the potential of various biological
materials found in nature is being researched in order to improve their properties for heavy metal
removal, learn about binding mechanisms and develop the most effective biosorbents for the
removal of contaminants, including heavy metals (Ibuot et al. 2020). Metal ions (typically in the
form of cations) are bound by cell membranes in a physicochemical procedure termed biosorption,
i.e., by negatively charged substances present in cell membranes. Understanding the biosorption
mechanisms that allow heavy metals to be removed is critical for process optimization (Irshad
et al. 2021). During sorption, numerous distinct mechanisms take place. Due to the complexity of
biological materials, several mechanisms may occur at different rates simultaneously. The following
mechanisms are covered by biosorption:
• Ion exchange is a reversible chemical reaction that occurs on materials containing relevant
functional groups and involves exchanging mobile ions for other ions of the same charge.
• Heavy metal ions bind to the cell membrane with functional elements, resulting in complexation.
• Physical adsorption is triggered by van der Waals forces, which are intermolecular interactions.
There is no chemical bonding in the case of physisorption.
Heavy metals are bound by cells that have ceased to function metabolically, which is the major
benefit of biosorption. As feeding living biomass necessitates an additional source of nutrients and
energy, contaminants can be removed by dead organisms, simplifying and lowering the cost of the
process (Jin et al. 2020). In this process, cell membranes play a crucial role. Before accessing the cell
membrane and cytoplasm, all metal ions pass through the cell wall. As a cell wall’s structure is made
up of diverse polysaccharides and proteins, there are several active places for metal ion binding.
Like commercially available resins, a cell wall can be treated as a complicated ion exchanger (Joshi
et al. 2011). Different microbial cell wall compositions and intercellular changes have a major impact
60 Bioremediation for Sustainable Environmental Cleanup
on the amount of adsorbed metal ions. Microbial cells have a surface made up of macromolecules
containing a lot of charged functional elements, such as:
• Phosphate
• Amino
• Carboxylic
• Hydroxyl
Due to the lack of carboxylic and phosphate acid residues, the cell surface is frequently
negatively charged. This enables passive cation binding on the cell surface. Positively charged metal
ions are drawn to the cell and adsorbed on the cell’s negatively charged surface (Kadukova 2016).
The entire process is passive and takes place without the cell’s metabolic processes. In Figure 4.2,
the entire cycle of wastewater treatment with biosorption is shown.
First, heavy metal effluent is combined with biomass. Metal ions adsorb on the surface of
bacteria as a result of this (biosorption). After that, biomass regeneration (desorption) is carried out,
and metals can be collected from the leftover liquid portion (Leong and Chang 2020). The following
are the phases of a typical biosorption laboratory procedure:
I. A solution sample of volume V containing heavy metals at concentration Ci is created in the
first phase.
II. In the second step, biomass M (g) is added to the sample under investigation.
III. To achieve equilibrium, the solution with biomass is vigorously agitated for about 16 hr.
IV. The biomass is then separated from the solution in the fourth phase (centrifugation, filtration).
V. The solution is then analyzed for metal ions after being treated with biosorption (Cf ).
Figure 4.2. Removal and recovery of heavy metals from wastewater by microbial biomass.
Bioaccumulation and Biosorption: The Prospects and Future Applications 61
The following formula (Eq. 4.1) is used to compute biosorption in mg metal per g biomass:
v(ci – cf)
q= Eq. 4.1
M
where:
q – number of adsorbed metals (mgg–1)
V – is the sample quantity of the solution (in millilitres).
Ci – metal ion concentration in solution at the start (mg L–1)
Cf – metal ion concentration in solution at the end (mg L–1)
M – mass of biomass (g)
4.5 Bioaccumulation
Different elements from the surrounding environment are absorbed and retained by living organisms.
Bioaccumulation is the process of harmful metals or organic compounds being bound inside a cell
structure. The biological activity of the biomass is critical in the bioaccumulation process (Loppi et
al. 2020). As a result of their metabolic functions, cells must be alive in order to adsorb pollutants.
The procedure should be kept under constant control in order to get the intended results. During
bioaccumulation, metal ions are absorbed by the entire cell. Metals enter the cells of living beings
through the same pathways as nutrients do. Heavy metals are absorbed by unicellular organisms
with the help of essential minerals like calcium and magnesium (Louati et al. 2020). Plants absorb
nutrients from water or soil through their roots, whereas mammals absorb nutrients through their
digestive or respiratory systems. There are two stages of bioaccumulation. Metal ions are bonded
on the cell surface in the first stage. The metabolically passive phase of the process is identical to
the biosorption mechanism. Metal ions are then transferred into the cell. It is possible only when the
cells are metabolically active in the second half of this process (Maal Bared 2020).
The quantity of biomass grows if correct conditions for organism growth are maintained in the
second stage. In comparison to biosorption, this allows (Nguyen et al. 2020) for the binding of larger
amounts of metal ions (Maal Bared 2020). Precipitation can lead to metal accumulation in microbial
biomass. It is critical, however, that precipitated metals were linked to the cell. Citrobacter, for
example, generates phosphatase, an acidic enzyme. Heavy metals are precipitated in the cell wall
in the form of barely soluble acidic phosphate when this enzyme is present. Denitrification caused
by Achromobactin dentifrices contributes to environmental acidity and hence increases metal
precipitation. Polysaccharides found in the cell walls of fungus and algae, such as cellulose, chitin
and alginates, also help to capture metals (Naik and Dubey 2013).
4.7.1 pH
As biosorption and bioaccumulation are analogous to ion exchange in some ways, the pH of a
solution has a big impact on heavy metal elimination. The amount of binding sites available on the
surface of cells is affected by pH (Deng et al. 2011). The accessible binding sites in a cell bind to
hydrogen cations in a solution when the pH is low. As a result, the number of accessible places is
limited, and fewer metal cations can be adsorbed. However, when pH rises, so does the amount of
active areas with a negative charge that attract cations (Devatha and Shivani 2020).
4.7.2 Temperature
Temperature affects the stability of metal ions in a solution as well as metal-cell complexes.
Temperatures between 20 and 35 degrees celsius, on the other hand, have little effect on biosorption
and bioaccumulation (Dey et al. 2020). Higher temperatures improve biomass sorption capability,
but they can also damage the sorption material.
which can disrupt the reactor’s balance (El Sayed and El Sayed 2020). The characteristics of cell
walls, which are critical for heavy metal adsorption, can be influenced by the age of the biomass.
The link between biomass age and heavy metal adsorption is not understood well, according to
several observations. Older cultures may have a broader ability to remove metals than younger
cultures, depending on the organisms used during the sorption process and biomagnification, or vice
versa. (Elahian et al. 2017).
References
Abdul Jaffar, A. H., M. Tamilselvi, A. S. Akram, M. L. Kaleem Arshan and V. Sivakumar. 2015. Comparative study
on bioremediation of heavy metals by solitary ascidian, Phallusia nigra, between Thoothukudi and Vizhinjam
ports of India. Ecotoxicol. Environ. Saf. 121: 93–99.
Ahmed, D. A. E., S. F. Gheda and G. A. Ismail. 2021. Efficacy of two seaweeds dry mass in bioremediation of
heavy metal polluted soil and growth of radish (Raphanus sativus L.) plant. Environ. Sci. Pollut. Res. 28:
12831–12846.
Al-Ansari, M. M., H. Benabdelkamel, R. H. AlMalki, A. M. Abdel Rahman, E. Alnahmi and A. Masood. 2021.
Effective removal of heavy metals from industrial effluent wastewater by a multi metal and drug resistant
Pseudomonas aeruginosa strain RA-14 using integrated sequencing batch reactor. Environ. Res. 199: 111240.
Arroyo-Herrera, I., B. Roman-Ponce, A. L. Resendiz-Martinez, S. P. Estrada-de Los, E. T. Wang and M. S. Vasquez-
Murrieta. 2021. Heavy-metal resistance mechanisms developed by bacteria from Lerma-Chapala basin. Arch.
Microbiol. 203: 1807–1823.
Bai, H., S. Wei, Z. Jiang, M. He, B. Ye and G. Liu. 2019. Pb (II) bioavailability to algae (Chlorella pyrenoidosa)
in relation to its complexation with humic acids of different molecular weight. Ecotoxicol. Environ. Saf.
167: 1–9.
Bai, L., H. Xu, C. Wang, J. Deng and H. Jiang. 2016. Extracellular polymeric substances facilitate the biosorption of
phenanthrene on cyanobacteria Microcystis aeruginosa. Chemosphere. 162: 172–180.
Baselga-Cervera, B., J. Romero-Lopez, C. Garcia-Balboa, E. Costas and V. Lopez-Rodas. 2018. Improvement of the
uranium sequestration ability of a Chlamydomonas sp. (ChlSP Strain) isolated from extreme uranium mine
tailings through selection for potential bioremediation Application. Front. Microbiol. 9: 523.
Bindschedler, S., T. Q. T. Vu Bouquet, D. Job, E. Joseph and P. Junier. 2017. Fungal biorecovery of gold from
E-waste. Adv. Appl. Microbiol. 99: 53–81.
Boriova, K., S. Cernansky, P. Matus, M. Bujdos, A. Simonovicova and M. Urik. 2019. Removal of aluminium from
aqueous solution by four wild-type strains of Aspergillus niger. Bioprocess. Biosyst. Eng. 42: 291–296.
Cachada, A., R. Pereira, E. F. da Silva and A. C. Duarte. 2014. The prediction of PAHs bioavailability in soils using
chemical methods: state of the art and future challenges. Sci. Total Environ. 472: 463–480.
Chamekh, A., O. Kharbech, R. Driss-Limam, C. Fersi, M. Khouatmeya and R. Chouari. 2021. Evidences for
antioxidant response and biosorption potential of Bacillus simplex strain 115 against lead. World J. Microbiol.
Biotechnol. 37: 44.
Bioaccumulation and Biosorption: The Prospects and Future Applications 65
Chen, S. H., Y. L. Cheow, S. L. Ng, S. L and A. S. Y. Ting. 2019. Mechanisms for metal removal established via
electron microscopy and spectroscopy: a case study on metal tolerant fungi Penicillium simplicissimum. J.
Hazard. Mater. 362: 394–402.
Chen, X., M. Zheng, G. Zhang, F. Li, H. Chen and Y. Leng. 2020. The nature of dissolved organic matter determines
the biosorption capacity of Cu by algae. Chemosphere 252: 126465.
Chojnacka, K. 2010. Biosorption and bioaccumulation—the prospects for practical applications. Environ. Int.
36: 299–307.
Danouche, M., G. N. El and A. H. El. 2021. Phycoremediation mechanisms of heavy metals using living green
microalgae: physicochemical and molecular approaches for enhancing selectivity and removal capacity.
Heliyon 7: e07609.
Deng, Z., L. Cao, H. Huang, X. Jiang, W. Wang, Y. Shi et al. 2011. Characterization of Cd- and Pb-resistant fungal
endophyte Mucor sp. CBRF59 isolated from rapes (Brassica chinensis) in a metal-contaminated soil. J.
Hazard. Mater. 185: 717–724.
Devanesan, S. and M. S. AlSalhi. 2021. Effective removal of Cd(2+), Zn(2+) by immobilizing the non-absorbent
active catalyst by packed bed column reactor for industrial wastewater treatment. Chemosphere 277: 130230.
Devatha, C. P. and S. Shivani. 2020. Novel application of maghemite nanoparticles coated bacteria for the removal of
cadmium from aqueous solution. J. Environ. Manage. 258: 110038.
Dey, P., A. Malik, A. Mishra, D. K. Singh, B. M. von and N. Jehmlich. 2020. Mechanistic insight to mycoremediation
potential of a metal resistant fungal strain for removal of hazardous metals from multimetal pesticide matrix.
Environ. Pollut. 262: 114255.
Du, J., P. Sun, Z. Feng, X. Zhang and Y. Zhao. 2016. The biosorption capacity of biochar for 4-bromodiphengl ether:
study of its kinetics, mechanism, and use as a carrier for immobilized bacteria. Environ. Sci. Pollut. Res.
23: 3770–3780.
El Sayed, M. T. and A. S. A. El-Sayed. 2020. Tolerance and mycoremediation of silver ions by Fusarium solani.
Heliyon. 6: e03866.
Elahian, F., S. Reiisi, A. Shahidi and S. A. Mirzaei. 2017. High-throughput bioaccumulation, biotransformation,
and production of silver and selenium nanoparticles using genetically engineered Pichia pastoris. Nanomed.
13: 853–861.
Fabre, E., M. Dias, M. Costa, B. Henriques, C. Vale, C. B. Lopes et al. 2020. Negligible effect of potentially toxic
elements and rare earth elements on mercury removal from contaminated waters by green, brown and red
living marine macroalgae. Sci. Total Environ. 72: 138133.
Fargasova, A., I. Ondrejkovicova, Z. Kramarova and Z. Faberova. 2010. Changes in physiological activity of
algae Desmodesmus quadricauda after active bioaccumulation of newly prepared and characterized Fe(III)
complexes with pyridine-3-carboxamide (pca) by living algal cells. Bioresour. Technol. 101: 6410–6416.
Flouty, R. and G. Estephane. 2012. Bioaccumulation and biosorption of copper and lead by a unicellular algae
Chlamydomonas reinhardtii in single and binary metal systems: a comparative study. J. Environ. Manage.
111: 106–114.
Gajda-Meissner, Z., K. Matyja, D. M. Brown, M. G. J. Hartl and T. F. Fernandes. 2020. Importance of surface coating
to accumulation dynamics and acute toxicity of copper nanomaterials and dissolved copper in Daphnia magna.
Environ. Toxicol. Chem. 39: 287–299.
Garg, S. K., M. Tripathi and T. Srinath. 2012. Strategies for chromium bioremediation of tannery effluent. Rev.
Environ. Contam. Toxicol. 217: 75–140.
Gerber, U., R. Hubner, A. Rossberg, E. Krawczyk-Barsch and M. L. Merroun. 2018. Metabolism-dependent
bioaccumulation of uranium by Rhodosporidium toruloides isolated from the flooding water of a former
uranium mine. PLoS. One. 13: e0201903.
Hansda, A., V. Kumar and A. Mishra. 2016. A comparative review towards potential of microbial cells for heavy metal
removal with emphasis on biosorption and bioaccumulation. World J. Microbiol. Biotechnol. 32: 170.
Hlihor, R. M., M. Diaconu, F. Leon, S. Curteanu, T. Tavares and M. Gavrilescu. 2015. Experimental analysis
and mathematical prediction of Cd(II) removal by biosorption using support vector machines and genetic
algorithms. N. Biotechnol. 32: 358–368.
Ibuot, A., R. E. Webster, L. E. Williams and J. K. Pittman. 2020. Increased metal tolerance and bioaccumulation
of zinc and cadmium in Chlamydomonas reinhardtii expressing a AtHMA4 C-terminal domain protein.
Biotechnol. Bioeng. 117: 2996–3005.
Irshad, S., Z. Xie, S. Mehmood A. Nawaz, A. Ditta and Q. Mahmood. 2021. Insights into conventional and recent
technologies for arsenic bioremediation: a systematic review. Environ. Sci. Pollut. Res. 28: 18870–18892.
Jin, C. S., R. J. Deng, B. Z. Ren, B. L. Hou and A. S. Hursthouse. 2020. Enhanced biosorption of Sb(III) onto
living Rhodotorula mucilaginosa strain DJHN070401: Optimization and Mechanism. Curr. Microbiol.
77: 2071–2083.
66 Bioremediation for Sustainable Environmental Cleanup
Joshi, P. K., A. Swarup, S. Maheshwari, R. Kumar and N. Singh. 2011. Bioremediation of heavy metals in liquid
media through fungi isolated from contaminated sources. Indian J. Microbiol. 51: 482–487.
Kadukova, J. 2016. Surface sorption and nanoparticle production as a silver detoxification mechanism of the
freshwater alga Parachlorella kessleri. Bioresour. Technol. 216: 406–413.
Kumar, P. A., Y. S. Tseng, S. R. Rani, C. W. Chen, J. S. Chang and D. C. Di. 2021. Novel application of microalgae
platform for biodesalination process: A review. Bioresour. Technol. 337: 125343.
Kumar, R., S. Singh and O. V. Singh. 2007. Bioremediation of radionuclides: emerging technologies. OMICS.
11: 295–304.
Leong, Y. K. and J. S. Chang. 2020. Bioremediation of heavy metals using microalgae: Recent advances and
mechanisms. Bioresour. Technol. 303: 122886.
Loppi, S., A. Vannini, F. Monaci, D. Dagodzo, F. Blind, M. Erler et al. 2020. Can chitin and chitosan replace the lichen
Evernia prunastri for environmental biomonitoring of Cu and Zn Air contamination? Biology. (Basel), 9.
Louati, I., J. Elloumi-Mseddi, W. Cheikhrouhou, B. Hadrich, M. Nasri, S. Aifa et al. 2020. Simultaneous cleanup of
Reactive Black 5 and cadmium by a desert soil bacterium. Ecotoxicol. Environ. Saf. 190: 110103.
Maal-Bared, R. 2020. Operational impacts of heavy metals on activated sludge systems: the need for improved
monitoring. Environ. Monit. Assess. 192: 560.
Naik, M. M. and S. K. Dubey. 2013. Lead resistant bacteria: lead resistance mechanisms, their applications in lead
bioremediation and biomonitoring. Ecotoxicol. Environ. Saf. 98: 1–7.
Nguyen, T. Q., V. Sesin, A. Kisiala and R. J. N. Emery. 2020. The role of phytohormones in enhancing metal
remediation capacity of algae. Bull. Environ. Contam. Toxicol. 105: 671–678.
Parsania, S., P. Mohammadi and M. R. Soudi. 2021. Biotransformation and removal of arsenic oxyanions by
Alishewanella agri PMS5 in biofilm and planktonic states. Chemosphere 284: 131336.
Pushkar, B., P. Sevak, S. Parab and N. Nilkanth. 2021. Chromium pollution and its bioremediation mechanisms in
bacteria: a review. J. Environ. Manage. 287: 112279.
Qiu, J., X. Song, S. Li, B. Zhu, Y. Chen, L. Zhang et al. 2021. Experimental and modeling studies of competitive
Pb (II) and Cd (II) bioaccumulation by Aspergillus niger. Appl. Microbiol. Biotechnol. 105: 6477–6488.
Reddy, K., N. Renuka, S. Kumari and F. Bux. 2021. Algae-mediated processes for the treatment of antiretroviral drugs
in wastewater: Prospects and challenges. Chemosphere 280: 130674.
Singh, S. and V. Kumar. 2020. Mercury detoxification by absorption, mercuric ion reductase, and exopolysaccharides:
a comprehensive study. Environ. Sci. Pollut. Res. 27: 27181–27201.
Thesai, A. S., G. Nagarajan, S. Rajakumar, A. Pugazhendhi and P. M. Ayyasamy. 2021. Bioaccumulation of fluoride
from aqueous system and genotoxicity study on Allium cepa using Bacillus licheniformis. J. Hazard. Mater.
407: 124367.
Yaashikaa, P. R., K. P. Senthil, S. Varjani and A. Saravanan. 2020. Rhizoremediation of Cu(II) ions from contaminated
soil using plant growth promoting bacteria: an outlook on pyrolysis conditions on plant residues for methylene
orange dye biosorption. Bioeng. 11: 175–187.
Ying T. H., D. Ishii, A. Mahara, S. Murakami, T. Yamaoka, K. Sudesh, R. Samian, M. Fujita, M. Maeda and
T. Iwata. 2008. Scaffolds from electrospun polyhydroxyalkanoate copolymers: fabrication, characterization,
bioabsorption and tissue response. Biomaterials. 29(10): 1307–17.
Chapter 5
PAHs in Terrestrial
Environment and their
Phytoremediation
Sandip Singh Bhatti,1 Astha Bhatia,2 Gulshan Bhagat,2
Simran Singh,2 Salwinder Singh Dhaliwal,3
Vivek Sharma,3 Vibha Verma,3 Rui Yin4 and
Jaswinder Singh5,*
5.1 Introduction
Polycyclic Aromatic Hydrocarbons (PAHs) are organic compounds that are present everywhere in
one’s surroundings. PAHs have high melting and boiling points which facilitates them to remain solid
at room temperature. Other properties of PAHs include low vapor pressure and very low aqueous
solubility which decrease with an increase in molecular weight. The increase in the molecular weight
of PAHs is also related to their resistance towards oxidation and reduction processes. According to
IUPAC (International Union of Pure and Applied Chemistry), phenanthrene and anthracene are the
simplest PAHs (Rengarajan et al. 2015).
1
Department of Chemistry, Lovely Professional University, Phagwara, 144001-India.
Email: [email protected], [email protected]
2
Department of Botanical and Environmental Sciences, Guru Nanak Dev University, Amritsar, 143005-India.
3
Department of Soil Science, Punjab Agricultural University, Ludhiana, Punjab, 141004, India.
4
Institute of Ecology, College of Urban and Environmental Sciences, and Key Laboratory for Earth Surface Processes of the
Ministry of Education, Peking University, 100871, Beijing, China.
5
Department of Zoology, Khalsa College, Amritsar, 143005-India.
* Corresponding author: [email protected]
68 Bioremediation for Sustainable Environmental Cleanup
PAHs released from the industries into the marine ecosystems are the main source of
contaminated marine ecosystems, whereas PAHs in the air are mainly from two sources which
include stationary (e.g., industrial production including power plant, coking plant, boiler and
waste incinerator and residential combustion like smoking, cooking, etc.) and mobile sources (e.g.,
exhausts from railways, motor vehicle, aircraft, etc.). PAHs in food (or the food cycle) is a major
health concern (Premnath et al. 2021). PAHs in the soil through the precipitation is carried to the
surface/groundwater and incorporated into crops, which after human consumption is accumulated
in human and other organisms via the food chain (Wang et al. 2018).
In terrestrial environment, PAHs are covered by the SOM (Soil Organic Matter) and soil
minerals due to their hydrophobic, lipophilic and semi volatile properties resulting in their decreased
degradation by microorganisms and finally their accumulation into the soil in large quantities
(Mazarji et al. 2021). Due to the carcinogenic toxicity of PAHs, there are strict regulations to limit
their release into the natural sources (Chatzimichail et al. 2021).
Out of 100 detected PAHs (parent and alkylated derivatives) 16 are listed in Clean Water
Act, 1972 (United States of America) which are presented in Table 5.1, whereas eight are listed
by European Commission (2013) due to their potential risk to mankind and ecological well-being
(Table 5.2) (Ofori et al. 2021).
Table 5.1. List of 16 PAHs under Clean Water Act, 1972 (USEPA 2014).
1. Chrysene 9. Napthalene
2. Acenaphthylene 10. Benzo[b]fluoranthene
3. Acenaphthene 11. Benzo[k]fluoranthene
4. Phenanthrene 12. Benzo[a]pyrene
5. Anthracene 13. Benzo[a]anthracene
6. Fluoranthene 14. Dibenz [a,h]anthracene
7. Fluorene 15. Benzo[g,h,i]perylene
8. Pyrene 16. Indo[1,2,3-cd] pyrene
Table 5.2. List of 8 PAHs under Water Framework Directive (European Commission 2013).
1. Napthalene 5. Anthracene
2. Benzo[b]fluoranthene 6. Fluoranthene
3. Benzo[k]fluoranthene 7. Indeno [1,2,3-cd]
4. Benzo[a]pyrene 8. Benzo[g,h,i]perylene
Based on the origin of production of PAHs, sources of PAHs can be categorized into three types,
i.e., pyrogenic, petrogenic and biogenic. Pyrogenic PAHs are produced when there is incomplete
combustion of fossil fuels at very high temperatures under anaerobic conditions. Petrogenic PAHs
are mainly present in crude and refined petroleum. These PAHs enter the soil during storage,
transport and leakage from oil refineries. Biogenic PAHs are produced by microorganisms, algae,
phytoplanktons and plants (Abdel-Shafy and Mansour 2016).
and
Water pollution through
surface run off from
polluted soil into Acts as
waterbodies immunotoxins
Effects of
PAH in soil
Bioaugmentation
Sr. No. Types of Soil Sites Sampling time/study Extraction methods No. PAHs Range of PAHs References
year
1 Air-port soil Indira Gandhi International November 2005–May Ultra-sonication method 12 2394–7529 Ray et al. 2008
Airport, Delhi, India 2006 (ng g–1)
2 Traffic soil Delhi, India Winter season, 2006 Ultra-sonication 16 1062–9652 Agarwal 2009
Method (µg kg–1)
3 Forest-fire region soil South Korea 2001 Accelerated Solvent 16 153–1,570 Kim et al. 2011
Extraction (ASE) (ng g–1)
4 Soil from brick Mexico (North America) - Soxhlet extraction 13 7–1384 (ng g–1) Barrán-Berdón et al.
manufacturing site 2012
(San Nicolás)
5 Urban soil Kurukshetra, June 2012 Ultra-sonication 16 19.1–2538.0 Kumar et al. 2013
India (µg kg–1)
71
...Table 5.3 contd.
72
Sr. No. Types of Soil Sites Sampling time/study Extraction methods No. PAHs Range of PAHs References
73
74 Bioremediation for Sustainable Environmental Cleanup
into less toxic forms under both aerobic as well as anaerobic conditions. In aerobic degradation, CO2
and water are formed as by-products, while in anaerobic conditions, methane is formed. Some of
the bacterial species involved in PAHs remediation are Pseudomonas, Rhodococcus, Bacillus and
Mycobacterium (Imam et al. 2022). Gamma-proteobacteria and actinobacteria are the two major
bacterial lineages that degrade PAHs in PAH-contaminated soils (Chaudhary et al. 2015). Apart
from bacteria, fungi are also employed for remediation known as mycoremediation due to their co
metabolic activity (Srivastava and Kumar 2019).
Phytoremediation is a broader term, it includes many techniques where plants are used to
remove, detoxify or immobilize contaminants like PAHs in soil. It is even more efficient, eco-friendly
and cost-effective than bioremediation. Phytoextraction, rhizoextraction, phytovolatilization,
phyto stabilization, phyto stimulation, phyto transformation, phyto assimilation, phyto reduction,
etc., are various green technologies of phytoremediation. Plants involved in the phytoremediation
process are called “hyperaccumulators” or “phytoremediators”. The process in which plants
remove contaminants from soil by storing them in tissues is called phytoextraction/rhizoextraction/
phytofilteration. In the phytovolatilization process, plants remove the contaminants through
volatilization. In phyto stabilization, plants reduce the mobility of contaminants. Phyto stabilization
followed by the addition of microbes degrade the immobile contaminants. The physical, chemical
or biological remediation methods can be employed both in-situ and ex-situ. In in-situ, remediation
is done at the original contaminated soil whereas in ex-situ, contaminated soil is transferred to
another location for remediation. Bionano-remediation is also an emerging technique for PAHs
degradation. The use of nanoparticles like Nano zero-valent iron and metal oxides, carbon-based
and polymer-based materials with biotechnological tools could be a promising approach for PAHs
remediation. It is observed that the addition of nanoparticles in soil positively affects the mobility
of PAHs, decreases the phytotoxicity and microbial flora remain unaffected (Mazarji et al. 2021).
Combinations of physical-biological, chemical-biological or all the three (physical-chemical
biological) known as integrated methods are also used as remediation techniques for PAHs (Patel
et al. 2020). Although various remediation technologies are available for PAH contaminated soils,
but out of them, the most suitable and sustainable technology is the phytoremediation (Dolatabadi
et al. 2021).
5.3.1.1 Phytoextraction
In phytoextraction, contaminants are absorbed by roots followed by their translocation and
accumulation in their aboveground biomass (Sreelal and Jayanthi 2017).
Screening of suitable plant species is the key and most straightforward strategy for successful
phytoextraction, i.e., the plant must be efficient in accumulating contaminants in the aerial parts.
Besides hyperaccumulation, the plant to act as eminently suitable for phytoextraction must
also possess traits like (1) rapid growth and production of large biomass; (2) vast root systems;
(3) easy cultivation and harvesting management; (4) preferably be repulsive to herbivores to avoid
the entrance in food chain (Seth 2012). However, natural hyperaccumulating plants lack these
characteristics thus limiting the phytoextraction potential (Chaney et al. 2005). To overcome the
problem, research has focused to modify or engineer large biomass producing non-hyperaccumulator
plants to achieve the above-mentioned attributes. To date, numerous hyperaccumulator plants ranging
from annual herbs to perennial shrubs and trees, have been used for phytoextraction. Phytoextraction
is considered advantageous as it does not alter the landscape, preserves the ecosystem and is cost-
effective, thus considered as the most commercially promising technique. However, several factors
such as lower bioavailability and absorption of metal in the roots limit the metal’s phytoextraction
by plants. However, the technique has been so far used for heavy metals (Jacobs et al. 2017, Guo
et al. 2020).
PAHs in Terrestrial Environment and their Phytoremediation 77
Phyto stabilization
on soil contaminated with polycyclic aromatic hydrocarbons was also studied. The concentration of
phenanthrene was reduced to greater extent with the integrated application of black locust and soil
amendments after 1 yr of plant growth. The uptake of trace elements Cu, As, Cd, Pb and Sb to leaves
was low (Wawra et al. 2018).
5.3.1.3 Phytovolatilization
Phytovolatilization has been observed for numerous contaminants including volatile organic
compounds and inorganic contaminants. Based on the possible mechanisms, the technique has
been divided into two categories, i.e., direct phytovolatilization by stems or leaves and indirect
phytovolatilization from the root zone. In direct phytovolatilization, the pollutants are absorbed by
roots, transported through the xylem and excreted to the atmosphere from the aerial plant parts in the
volatile form. However, in indirect phytovolatilization process, volatile organic contaminants flux
increases from the subsurface due to plant roots activities, e.g., increasing soil permeability, chemical
transport via hydraulic redistribution or water table fluctuations (Limmer and Burken 2016). This
strategy may prove advantageous as the contaminant can be transformed from more a toxic form to
a relatively lesser toxic substance, however, there might be the possibility that the resultant form is
still potentially toxic, and thus resettle into the environment. The phytovolatilization can be applied
for contaminants present in soil, sediment or water, particularly, for organic contaminants. Till date,
this technique has been used for heavy metals (Shrestha et al. 2006, Sakakibara et al. 2010).
5.3.1.4 Phytodegradation
‘Phytodegradation’ is a subset technique of phytoremediation under which organic contaminants are
absorbed by the roots followed by degradation due to catalytic action of enzymes which are involved
in metabolism of the plant (Newman and Reynolds 2004, Sharma and Pandey 2014). The enzymes
involved in the phytodegradation process are dehalogenase, peroxidase, nitroreductase, nitrilase and
phosphatase (Chatterjee et al. 2013). Phytodegradation is primarily used for the remediation sites
by contamination with organic pollutants like and PAHs (Muthusaravanan et al. 2018). To date,
numerous plants have been used for phytodegradation of organic contaminants. Lupinus luteus in
association with endophytic bacteria has shown immense phytodegradation potential in landfill soils
of the Iberian Peninsula contaminated with PAHs (Gutiérrez-Ginés et al. 2014). He and Chi (2019)
investigated the phytodegradation of phenanthrene and pyrene using aquatic plants, Vallisneria
spiralis and Hydrilla verticillata, in PAH-polluted sediments. Among them, sediments planted with
V. spiralis showed the highest dissipation of phenanthrene and pyrene (85.9 and 79.1%) as compared
to sediments planted with H. verticillata and unplanted sediment. Phytodegradation of phenanthrene
and pyrene using the maize plant has been confirmed using GC-MS analysis. The degradation rate
of phenanthrene was found to be faster than that of pyrene and prominently occurred in the roots
(Houshani et al. 2021).
5.3.1.5 Rhizofiltration
‘Rhizofiltration’or ‘phytofiltration’refers to the process that uses both terrestrial and aquatic plants in
adsorbing contaminants in the surrounding root zone (rhizosphere), concentrating and precipitating
them on or within the root (Rahman and Hasegawa 2011). The plants’ roots are harvested after
becoming saturated. The plant species which possess high tolerance towards metal toxicity and
have a high surface area for absorption of metal are preferred (e.g., Salix spp., Populus spp.,
Brassica spp.). Terrestrial plants have been reported as a suitable candidate for rhizofilteration as
they possess a more developed and fibrous root structure, thus provide a higher surface area for
absorption of contaminants (Cule et al. 2016). This strategy can be applied for the remediation of
contaminated sites loaded with heavy metals and radionuclides. The factors that limit the application
of this technique are pH adjustment, hydroponic cultivation in a greenhouse, frequent harvests and
proper disposal of the plants (Kristanti et al. 2020).
PAHs in Terrestrial Environment and their Phytoremediation 79
Research studies demonstrated that sunflowers have shown greater potential for phytoremediation
of soil contaminated with radionuclides (caesium and strontium) in Chernobyl, Ukraine. The
results indicated that Cs was accumulated in roots, whereas Sr uptake was recorded in the shoots
(Prasad 2007). Another study reported that rhizofiltration using and Phaseolus vulgaris has shown
remediation of uranium contaminated groundwater (Lee and Yang 2010). The aromatic medicinal
plant, Plectranthus amboinicus, has been evaluated as a candidate for rhizofilteration of lead-
containing wastewater. The results indicated that P. amboinicus showed potential tolerance towards
Pb toxicity and accumulates Pb maximum in the roots, whereas translocation in the stem and leaf was
limited (Ignatius et al. 2014). In hydroponic experiments, Typha angustifolia and Acorus calamus
(aquatic) and Pandanus amaryllifolius (terrestrial) plants were evaluated for rhizofiltration of Cd
and Zn. Among these, T. angustifolia showed maximum tolerance towards HM toxicity without loss
in dry biomass production. With maximum HM accumulation in roots, T. angustifolia was found to
be the most suitable candidate plant for phytoremediation in constructed wetlands and aquatic plant
systems (Woraharn et al. 2021).
5.3.1.6 Rhizodegradation
Rhizodegradation is an emerging technique for remediation of contaminated soil that involves
plant roots, plant-supplied nutrients and soil microorganisms. This process occurs through the
plant-supplied substrates such as bacteria, fungi and yeasts which favors the growth of microbial
communities in the rhizosphere to break down organic pollutants (Allamin et al. 2020, Rajkumari
et al. 2021). The factors affecting the rhizodegradation efficiency of processes are the type
of contaminant, the ability of microflora to degrade the contaminants and the bioavailability of
pollutants (Noroozi et al. 2017, Patel and Patra 2017). Plant roots provide air to the soil and release
exoenzymes and nutrients through root exudates. The advantages of rhizodegradation are complete
mineralization of the contaminant, lesser translocation of the contaminant to other plant parts or in
the atmosphere, low installation and maintenance costs (Dos Santos and Maranho 2018). However,
there exist some disadvantages of the process as it is a slow process and effective only on the surface
of contamination (20–25 cm of depth); the limited depth of the roots and requirement of fertilizers
by plants.
Various studies have reported the effective remediation of soils via rhizodegradation
contaminated including PAHs, pesticides and solvents containing benzene ring. Plants such as
Avicennia marina (Jia et al. 2016) or Lolium multiflorum (Hussain et al. 2022) have shown immense
potential towards the degradation of PAH following the rhizodegradation mechanism. The efficiency
of Rhizophora mangle to extract PAHs in mangrove sediment contaminated with crude oil has also
been studied. The soil planted with Rhizophora mangle showed higher removal of 16 PAHs as
compared to natural attenuation (Verâne et al. 2020).
from the contaminant type, the concentration of the contaminant should also be estimated prior
to the phytoremediation to prevent the toxicity effects on healthy plants. The non-aqueous phase
pollutants create adverse effects on plant growth. Pollutants with low bioavailability are difficult to
extract with phytoremediation process (EPA 2000).
References
Abdel-Shafy, H. I. and M. S. Mansour. 2016. A review on polycyclic aromatic hydrocarbons: source, environmental
impact, effect on human health and remediation. Egypt. J. Pet. 25(1): 107–123.
Agarwal, T. 2009. Concentration level, pattern and toxic potential of PAHs in traffic soil of Delhi, India. J. Hazard.
Mater. 171(1-3): 894–900.
PAHs in Terrestrial Environment and their Phytoremediation 81
Ali, H., E. Khan and M. A. Sajad. 2013. Phytoremediation of heavy metals—concepts and
applications. Chemosphere. 91(7): 869–881.
Allamin, I. A., M. I. E. Halmi, N. A. Yasid, S. A. Ahmad, S. R. S. Abdullah and Y. Shukor. 2020. Rhizodegradation
of petroleum oily sludge-contaminated soil using Cajanus cajan increases the diversity of soil microbial
community. Sci. Rep. 10(1): 1–11.
Ambade, B., S. S. Sethi and M. R. Chintalacheruvu. 2022. Distribution, risk assessment, and source apportionment
of polycyclic aromatic hydrocarbons (PAHs) using positive matrix factorization (PMF) in urban soils of East
India. Environ. Geochem. Health 1–15. https://doi.org/10.1007/s10653-022-01223-x.
Aralu, C. C., P. A. C. Okoye, K. G. Akpomie, H. O. Chukwuemeka‐Okorie and H. O. Abugu. 2022. Polycyclic
aromatic hydrocarbons in soil situated around solid waste dumpsite in Awka, Nigeria. Toxin Rev. 1–10. https://
doi.org/10.1080/15569543.2021.2022700.
Ashraf, S., Q. Ali, Z. A. Zahir, S. Ashraf and H. N. Asghar. 2019. Phytoremediation: Environmentally sustainable way
for reclamation of heavy metal polluted soils. Ecotoxicol. Environ. Saf. 174: 714–727.
Barrán-Berdón, A. L., V. Garcia Gonzalez, G. Pedraza Aboytes, I. Rodea-Palomares, A. Carrillo-Chavez, H. Gomez-
Ruiz and B. Verduzco Cuellar. 2012. Polycyclic aromatic hydrocarbons in soils from a brick manufacturing
location in central Mexico. Rev. Int. Contam. Ambient. 28(4): 277–288.
Bose, V. G., K. S. Shreenidhi and J. A. Malik. 2022. Phytoremediation of PAH-contaminated areas. pp. 141–156.
In: Advances in Bioremediation and Phytoremediation for Sustainable Soil Management. Springer, Cham.
Cetin, B., S. Yurdakul, M. Keles, I. Celik, F. Ozturk and C. Dogan. 2017a. Atmospheric concentrations,
distributions and air-soil exchange tendencies of PAHs and PCBs in a heavily industrialized area in Kocaeli,
Turkey. Chemosphere. 183: 69–79.
Cetin, B., F. Ozturk, M. Keles and S. Yurdakul. 2017b. PAHs and PCBs in an Eastern mediterranean megacity,
Istanbul: their spatial and temporal distributions, air-soil exchange and toxicological effects. Environ.
Pollut. 220: 1322–1332.
Chaney, R. L., J. S. Angle, M. S. McIntosh, R. D. Reeves, Y. M. Li, E. P. Brewer, K. Y. Chen, R. J. Roseberg,
H. Perner, E. C. Synkowski and C. L. Broadhurst. 2005. Using hyperaccumulator plants to phytoextract soil
Ni and Cd. Z. Naturforsch C 60(3-4): 190–198.
Chatterjee, S., A. Mitra, S. Datta and V. Veer. 2013. Phytoremediation protocols: An overview. In: D. Gupta (eds.).
Plant-Based Remediation Processes. Soil Biol. 35. https://doi.org/10.1007/978-3-642-35564-6_1.
Chatzimichail, S., F. Rahimi, A. Saifuddin, A. J. Surman, S. D. Taylor-Robinson and A. Salehi-Reyhani. 2021.
Hand-portable HPLC with broadband spectral detection enables analysis of complex polycyclic aromatic
hydrocarbon mixtures. Commun. Chem. 4(1): 1–14.
Chaudhary, P., H. Sahay, R. Sharma, A. K. Pandey, S. B. Singh, A. K. Saxena and L. Nain. 2015. Identification and
analysis of polyaromatic hydrocarbons (PAHs)—biodegrading bacterial strains from refinery soil of India.
Environ. Monit. Assess. 187(6): 1–9.
Cule, N., D. Vilotic, M. Nesic, M. Veselinovic, D. Drazic and S. Mitovic. 2016. Phytoremediation potential of Canna
indica L. in water contaminated with lead. Fresenius Environ. Bull. 25(11): 3728–3733.
Cunningham, S. D. and D. W. Ow. 1996. Promises and prospects of phytoremediation. Plant Physiol. 110(3): 715–719.
Dal Corso, G., E. Fasani, A. Manara, G. Visioli and A. Furini. 2019. Heavy metal pollutions: state of the art and
innovation in phytoremediation. Int. J. Mol. Sci. 20(14): 3412.
De Boer, J. and M. Wagelmans. 2016. Polycyclic aromatic hydrocarbons in soil–practical options for remediation.
CLEAN–Soil Air Water. 44(6): 648–653.
Deka, J., N. Baul, P. Bharali, K. P. Sarma and R. R. Hoque. 2020. Soil PAHs against varied land use of a small
city (Tezpur) of middle Brahmaputra Valley: Seasonality, sources, and long-range transport. Environ. Monit.
Assess. 192: 357.
Devi, N. L., I. C. Yadav, Q. Shihua, Y. Dan, G. Zhang and P. Raha. 2016. Environmental carcinogenic polycyclic
aromatic hydrocarbons in soil from Himalayas, India: Implications for spatial distribution, sources
apportionment and risk assessment. Chemosphere. 144: 493–502.
Dhaliwal, S. S., J. Singh, P. K. Taneja and A. Mandal. 2020. Remediation techniques for removal of heavy metals
from the soil contaminated through different sources: a review. Environ. Sci. Pollut. Res. 27(2): 1319–1333.
Dolatabadi, N., S. Mohammadi Alagoz, B. Asgari Lajayer and E. D. van Hullebusch. 2021. Phytoremediation of
polycyclic aromatic hydrocarbons-contaminated soils. pp. 419–445. In: Climate Change and the Microbiome.
Springer, Cham.
Dos Santos, J. J. and L. T. Maranho. 2018. Rhizospheric microorganisms as a solution for the recovery of soils
contaminated by petroleum: a review. J. Environ. Manag. 210: 104–113.
Dushenkov, V., P. N. Kumar, H. Motto and I. Raskin. 1995. Rhizofiltration: the use of plants to remove heavy metals
from aqueous streams. Environ. Sci. Technol. 29(5): 1239–1245.
82 Bioremediation for Sustainable Environmental Cleanup
Ke, C. L., Y. G. Gu and Q. Liu. 2017. Polycyclic aromatic hydrocarbons (PAHs) in exposed-lawn soils from 28
urban parks in the megacity Guangzhou: occurrence, sources, and human health implications. Arch. Environ.
Contam. Toxicol. 72(4): 496–504.
Kim, E. J., S. D. Choi and Y. S. Chang. 2011. Levels and patterns of polycyclic aromatic hydrocarbons (PAHs) in soils
after forest fires in South Korea. Environ. Sci. Pollut. Res. 18(9): 1508–1517.
Kotoky, R. and P. Pandey. 2020. Rhizosphere mediated biodegradation of benzo(A)pyrene by surfactin producing soil
bacilli applied through Melia azadirachta rhizosphere. Int. J. Phytoremediat. 22(4): 363–372.
Kristanti, R. A. W. J. Ngu, A. Yuniarto and T. Hadibarata. 2021. Rhizofiltration for removal of inorganic and organic
pollutants in groundwater: a review. Biointerafce Res. Appl. Chem. 4: 12326–12347.
Kumar, A. V., N. C. Kothiyal, S. Kumari, R. Mehra, A. Parkash, R. R. Sinha, S. K. Tayagi and R. Gaba. 2014.
Determination of some carcinogenic PAHs with toxic equivalency factor along roadside soil within a fast
developing northern city of India. J. Earth Syst. Sci. 123(3): 479–489.
Kumar, B., V. K. Verma, S. Kumar and C. S. Sharma. 2013. Probabilistic health risk assessment of polycyclic
aromatic hydrocarbons and polychlorinated biphenyls in urban soils from a tropical city of India. J. Environ.
Sci. Health, Part A 48(10): 1253–1263.
Kwon, H. O. and S. D. Choi. 2014. Polycyclic aromatic hydrocarbons (PAHs) in soils from a multi-industrial city,
South Korea. Sci. Total Environ. 470: 1494–1501.
Lee, M. and M. Yang. 2010. Rhizofiltration using sunflower (Helianthus annuus L.) and bean (Phaseolus vulgaris L.
var. vulgaris) to remediate uranium contaminated groundwater. J. Hazard. Mater. 173(1-3): 589–596.
Limmer, M. and J. Burken. 2016. Phytovolatilization of organic contaminants. Environ. Sci. Technol. 50(13): 6632–
6643.
Liu, J., S. Zhang, J. Jia, M. Lou, X. Li, S. Zhao, W. Chen, B. Xiao and Y. Yu. 2022. Distribution and source
apportionment of polycyclic aromatic hydrocarbons in soils at different distances and depths around three
power plants in Bijie, Guizhou Province. Polycycl. Aromat. Compd. https://doi.org/10.1080/10406638.2022
.2039232.
Lu, H., Y. Zhang, B. Liu, J. Liu, J. Ye and C. Yan. 2011. Rhizodegradation gradients of phenanthrene and pyrene in
sediment of mangrove (Kandelia candel (L.) Druce). J. Hazard. Mater. 196: 263–269.
Mazarji, M., T. Minkina, S. Sushkova, S. Mandzhieva, G. N. Bidhendi, A. Barakhov and A. Bhatnagar. 2021. Effect of
nanomaterials on remediation of polycyclic aromatic hydrocarbons-contaminated soils: a review. J. Environ.
Manag. 284: 112023.
Mudgal, V., N. Madaan and A. Mudgal. 2010. Heavy metals in plants: phytoremediation: plants used to remediate
heavy metal pollution. Agric. Biol. J. North Am. 1(1): 40–46.
Muthusaravanan, S., N. Sivarajasekar, J. S. Vivek, T. Paramasivan, M. Naushad, J. Prakashmaran, V. Gayathri and
O. K. Al-Duaij. 2018. Phytoremediation of heavy metals: mechanisms, methods and enhancements. Environ.
Chem. Lett. 16(4): 1339–1359.
Nayak, Y., S. Chakradhari, K. S. Patel, R. K. Patel, S. Yurdakul, H. Saathoff and P. Martín-Ramos. 2022. Distribution,
variations, fate and sources of polycyclic aromatic hydrocarbons and carbon in particulate matter, road dust,
and sediments in central India. Polycycl. Aromat. Compd. https://doi.org/10.1080/10406638.2022.2026991.
Nedjimi, B. 2021. Phytoremediation: a sustainable environmental technology for heavy metals decontamination. SN
Appl. Sci. 3: 286. https://doi.org/10.1007/s42452-021-04301-4.
Newman, L. A. and C. M. Reynolds. 2004. Phytodegradation of organic compounds. Curr. Opin. Biotechnol. 15(3):
225–230.
Nguyen, T. C., P. Loganathan, T. V. Nguyen, S. Vigneswaran, J. Kandasamy, D. Slee, G. Stevenson and R. Naidu.
2014. Polycyclic aromatic hydrocarbons in road-deposited sediments, water sediments, and soils in Sydney,
Australia: comparisons of concentration distribution, sources and potential toxicity. Ecotoxicol. Environ. Saf.
104: 339–348.
Noroozi, M., M. A. Amozegar, R. Rahimi, S. A. Shahzadeh Fazeli and G. Bakhshi Khaniki. 2017. The isolation and
preliminary characterization of native cyanobacterial and microalgal strains from lagoons contaminated with
petroleum oil in Khark Island. Biological Journal of Microorganism 5(20): 33–41.
Ofori, S. A., S. J. Cobbina, A. Z. Imoro, D. A. Doke and T. Gaiser. 2021. Polycyclic Aromatic Hydrocarbon
(PAH) pollution and its associated human health risks in the Niger delta region of Nigeria: A Systematic
Review. Environ. Process. 8(2): 455–482.
Patel, A. and D. D. Patra. 2017. A sustainable approach to clean contaminated land using terrestrial grasses.
pp. 305–331. In: Phytoremediation Potential of Bioenergy Plants. Springer, Singapore.
Patel, A. B., S. Shaikh, K. R. Jain, C. Desai and D. Madamwar. 2020. Polycyclic aromatic hydrocarbons: sources,
toxicity, and remediation approaches. Front. Microbiol. 11: 562813.
84 Bioremediation for Sustainable Environmental Cleanup
Phan Thi, L. A., N. T. Ngoc, N. T. Quynh, N. V. Thanh, T. T. Kim, D. H. Anh and P. H. Viet. 2020. Polycyclic aromatic
hydrocarbons (PAHs) in dry tea leaves and tea infusions in Vietnam: contamination levels and dietary risk
assessment. Environ. Geochem. Health. 42(9): 2853–2863.
Prasad, M. N. V. 2007. Sunflower (Helinathus annuus L.)—a potential crop for environmental industry/girasol
(Helianthus annuus L.)-cultivo potencial para la industria ecológica/le tournesol (Helianthus annuus L.)
culture potentielle dans l’industrie écologique. Helia. 30(46): 167–174.
Premnath, N., K. Mohanrasu, R. G. R. Rao, G. H. Dinesh, G. S. Prakash, V. Ananthi, K. Ponnuchamy, G. Muthusamy
and A. Arun. 2021. A crucial review on polycyclic aromatic Hydrocarbons-Environmental occurrence and
strategies for microbial degradation. Chemosphere. 280: 130608.
Qu, C., S. Albanese, A. Lima, D. Hope, P. Pond, A. Fortelli, N. Romano, P. Cerino, A. Pizzolante and B. De Vivo.
2019. The occurrence of OCPs, PCBs, and PAHs in the soil, air, and bulk deposition of the Naples metropolitan
area, southern Italy: implications for sources and environmental processes. Environ. Int. 124: 89–97.
Radziemska, M., M. D. Vaverková and A. Baryła. 2017. Phytostabilization—management strategy for stabilizing
trace elements in contaminated soils. Int. J. Environ. Res. Public Health. 14(9): 958.
Rahman, M. A. and H. Hasegawa. 2011. Aquatic arsenic: phytoremediation using floating
macrophytes. Chemosphere. 83(5): 633–646.
Rajan, S., K. R. Rex, M. Pasupuleti, J. Muñoz-Arnanz, B. Jiménez and P. Chakraborty. 2021. Soil concentrations,
compositional profiles, sources and bioavailability of polychlorinated dibenzo dioxins/furans, polychlorinated
biphenyls and polycyclic aromatic hydrocarbons in open municipal dumpsites of Chennai city, India. Waste
Manag. 131: 331–340.
Rajkumari, J., Y. Choudhury, K. Bhattacharjee and P. Pandey. 2021. Rhizodegradation of pyrene by a non
pathogenic Klebsiella pneumoniae isolate applied with Tagetes erecta L. and changes in the rhizobacterial
community. Front. Microbiol. 12: 593023.
Ray, S., P. S. Khillare, T. Agarwal and V. Shridhar. 2008. Assessment of PAHs in soil around the international airport
in Delhi, India. J. Hazard. Mater. 156(1-3): 9–16.
Ren, G., W. Ren, Y. Teng and Z. Li. 2015. Evident bacterial community changes but only slight degradation when
polluted with pyrene in a red soil. Front. Microbiol. 6: 22.
Rengarajan, T., P. Rajendran, N. Nandakumar, B. Lokeshkumar, P. Rajendran and I. Nishigaki. 2015. Exposure to
polycyclic aromatic hydrocarbons with special focus on cancer. Asian Pac. J. Trop. Biomed. 5(3): 182–189.
Sakakibara, M., A. Watanabe, M. Inoue, S. Sano and T. Kaise. 2010. Phytoextraction and phytovolatilization of arsenic
from As-contaminated soils by Pteris vittata. Vol. 121, p. 26. In: Proceedings of the Annual International
Conference on Soils, Sediments, Water And Energy.
Salehi, N., A. Azhdarpoor and M. Shirdarreh. 2020. The effect of different levels of leachate on phytoremediation
of pyrene-contaminated soil and simultaneous extraction of lead and cadmium. Chemosphere. 246: 125845.
Schwab, A. P. 1998. Phytoremediation of soils contaminated with PAHs and other petroleum compounds. In: Beneficial
effects of vegetation in contaminated soils workshop, Kansas State University, Manhattan, KS.
Seth, C. S. 2012. A review on mechanisms of plant tolerance and role of transgenic plants in environmental clean
up. Bot. Rev. 78(1): 32–62.
Sharma, P. and S. Pandey. 2014. Status of phytoremediation in world scenario. Int. J. Environ. Bioremediat.
Biodegrad. 2(4): 178–191.
Shrestha, B., S. Lipe, K. A. Johnson, T. Q. Zhang, W. Retzlaff and Z. Q. Lin. 2006. Soil hydraulic manipulation and
organic amendment for the enhancement of selenium volatilization in a soil–pickleweed system. Plant and
Soil. 288(1): 189–196.
Shukla, S., R. Khan, P. Bhattacharya, S. Devanesan and M. S. AlSalhi. 2022. Concentration, source apportionment and
potential carcinogenic risks of polycyclic aromatic hydrocarbons (PAHs) in roadside soils. Chemosphere. 292:
133413.
Singh, S. K. and A. K. Haritash. 2019. Polycyclic aromatic hydrocarbons: soil pollution and remediation. Int. J. Sci.
Environ. Technol. 16(10): 6489–6512.
Sreelal, G. and R. Jayanthi. 2017. Review on phytoremediation technology for removal of soil contaminant. Indian
J. Sci. Res. 14(1): 127–130.
Srivastava, S. and M. Kumar. 2019. Biodegradation of polycyclic aromatic hydrocarbons (PAHs): a sustainable
approach. pp. 111–139. In: Sustainable Green Technologies for Environmental Management. Springer,
Singapore.
Suman, S., A. Sinha and A. Tarafdar. 2016. Polycyclic aromatic hydrocarbons (PAHs) concentration levels,
pattern, source identification and soil toxicity assessment in urban traffic soil of Dhanbad, India. Sci. Total
Environ. 545: 353–360.
Tarafdar, A. and A. Sinha. 2018. Public health risk assessment with bioaccessibility considerations for soil PAHs at
oil refinery vicinity areas in India. Sci. Total Environ. 616: 1477–1484.
PAHs in Terrestrial Environment and their Phytoremediation 85
Tarafdar, A. and A. Sinha. 2019. Health risk assessment and source study of PAHs from roadside soil dust of a heavy
mining area in India. Arch. Environ. Occup. Health. 74(5): 252–262.
Tsibart, A. S. and A. N. Gennadiev. 2013. Polycyclic aromatic hydrocarbons in soils: sources, behavior, and indication
significance (a review). Eurasian Soil Sci. 46(7): 728–741.
USEPA. 2005. Guidelines for carcinogen risk assessment. EPA/630/P-03/001F2005: risk assessment forum. US
Environmental Protection Agency, Washington, DC.
USEPA. 2014. Priority pollutant list. https://www.epa.gov/sites/production/files/2015-09/documents/priority
pollutant-list-epa.pdf. Accessed on April 08, 2022.
Verâne, J., N. C. Dos Santos, V. L. da Silva, M. de Almeida, O. M. de Oliveira and Í. T. Moreira. 2020. Phytoremediation
of polycyclic aromatic hydrocarbons (PAHs) in mangrove sediments using Rhizophora mangle. Mar. Pollut.
Bull. 160: 111687.
Wan, X., M. Lei and T. Chen. 2016. Cost-benefit calculation of phytoremediation technology for heavy-metal
contaminated soil. Sci. Total Environ. 563: 796–802.
Wang, D., J. Ma, H. Li and X. Zhang. 2018. Concentration and potential ecological risk of PAHs in different layers of
soil in the petroleum-contaminated areas of the Loess Plateau, China. Int. J. Environ. Res. Public Health. 15(8):
1785.
Wang, J., X. Zhang, W. Ling, R. Liu, J. Liu, F. Kang and Y. Gao. 2017. Contamination and health risk assessment
of PAHs in soils and crops in industrial areas of the Yangtze River Delta region, China. Chemosphere. 168:
976–987.
Wang, R., G. Liu, C. L. Chou, J. Liu and J. Zhang. 2010. Environmental assessment of PAHs in soils around the Anhui
Coal District, China. Arch. Environ. Contam. Toxicol. 59(1): 62–70.
Wawra, A., W. Friesl-Hanl, M. Puschenreiter, G. Soja, T. Reichenauer, C. Roithner and A. Watzinger. 2018. Degradation
of polycyclic aromatic hydrocarbons in a mixed contaminated soil supported by phytostabilisation, organic
and inorganic soil additives. Sci. Total Environ. 628: 1287–1295.
Woraharn, S., W. Meeinkuirt, T. Phusantisampan and P. Chayapan. 2021. Rhizofiltration of cadmium and zinc in
hydroponic systems. Water Air Soil Pollut. 232(5): 1–17.
Yan, A., Y. Wang, S. N. Tan, M. L. Mohd Yusof, S. Ghosh and Z. Chen. 2020. Phytoremediation: a promising approach
for revegetation of heavy metal-polluted land. Front. Plant Sci. 11: 359.
Yuan, H., T. Li, X. Ding, G. Zhao and S. Ye. 2014. Distribution, sources and potential toxicological significance
of polycyclic aromatic hydrocarbons (PAHs) in surface soils of the Yellow River Delta, China. Mar. Pollut.
Bull. 83(1): 258–264.
Zamani, J., M. A. Hajabbasi, M. R. Mosaddeghi, M. Soleimani, M. Shirvani and R. Schulin. 2018. Experimentation
on degradation of petroleum in contaminated soils in the root zone of maize (Zea Mays L.) inoculated with
Piriformospora indica. Soil Sediment Contam. 27: 13–30.
Zgorelec, Z., N. Bilandzija, K. Knez, M. Galic and S. Zuzul. 2020. Cadmium and mercury phytostabilization from
soil using Miscanthus giganteus. Sci. Rep. 10(1): 1–10.
Zhang, S., H. Yao, Y. Lu, X. Yu, J. Wang, S. Sun, M. Liu, D. Li, Y.F. Li and D. Zhang. 2017. Uptake and translocation
of polycyclic aromatic hydrocarbons (PAHs) and heavy metals by maize from soil irrigated with wastewater.
Sci. Rep. 7(1): 1–11.
Zhao, L., H. Hou, Y. Shangguan, B. Cheng, Y. Xu, R. Zhao, Y. Zhang, X. Hua, X. Huo and X. Zhao. 2014. Occurrence,
sources, and potential human health risks of polycyclic aromatic hydrocarbons in agricultural soils of the coal
production area surrounding Xinzhou, China. Ecotoxicol. Environ. Saf. 108: 120–128.
Chapter 6
Fungal Strategies for the
Remediation of Polycyclic
Aromatic Hydrocarbons
Nitu Gupta,1 Sandipan Banerjee,2 Apurba Koley,3
Aman Basu,4 Nayanmoni Gogoi,1 Raza Rafiqul Hoque,1
Narayan Chandra Mandal 2 and
Srinivasan Balachandran 3,*
6.1 Introduction
A progressive increase in industrialization over several decades has caused the release of various
potentially hazardous contaminants such as xenobiotic compounds and heavy metals, that severely
affect the environment, i.e., soil, air and water. In these conditions, nowadays, environmental
pollutants are becoming a global concern for our ecosphere. These contaminants exhibit versatile
harmful effects on the human body and ecosystem. The significant human health hazards associated
with Polycyclic Aromatic Hydrocarbons (PAHs) include carcinogenic, mutagenic, teratogenic and
immune suppressants. It can rapidly assimilate in the gastrointestinal tract of mammals because
of its lipid-dissolvable nature (Abdel-Shafy and Mansour 2016). Such toxic substances can be
bioaccumulated in the food chain due to their non-polar, hydrophobic and lipophilic nature (Masih
et al. 2012). Several findings reveal that PAHs induce tumors in humans. This carcinogenic
characteristic of PAHs is co-relatable with their molecular structure. Structurally, carcinogenic
PAHs have a bay or K locale, showing a high affinity towards mammalian DNA. As a result,
PAHs-linked DNA adducts transformed the typical cell into a tumorigenic cell (Weis et al. 1998).
Hence, more attention should be paid to the remediation and restoration approaches for cleaning
up the contaminated environment from these pollutants. In this context, PAHs are considered one
of the prime concerns for the ecological malaise. PAHs are characterized as hydrophobic organic
compounds that are chemically constituted by two or more benzenoid rings. They are omnipresent,
1
Department of Environmental Science, Tezpur University, Napaam, Tezpur, Assam, 784028, India.
2
Mycology and Plant Pathology Laboratory, Department of Botany, Visva-Bharati, Santiniketan 731235, West Bengal,
India.
3
Department of Environmental Studies, Visva-Bharati, Santiniketan-731235, West Bengal, India.
4
Department of Biology, York University, Canada.
* Corresponding author: [email protected]
Fungal Strategies for the Remediation of Polycyclic Aromatic Hydrocarbons 87
and their concentration is elevated in the vicinity of the industries associated with petroleum and gas
production. On the other hand, anthropogenic combustion is the key contributor to PAHs pollution
(Banerjee and Mandal 2020, Ghosal et al. 2016).
PAHs compounds constitute two or more aromatic rings structurally arranged linearly, clusterly
or angularly. Generally, PAH compounds are comprised of carbon-hydrogen atoms; additionally,
nitrogen, oxygen and sulfur atoms can also be involved in forming heterocyclic aromatic compounds.
Broadly PAHs can be categorized into two groups: Low-Molecular-Weight PAHs (LMW-PAHs)
having less than four benzenoid rings whereas High-Molecular-Weight PAHs (HMW-PAHs) with
more than four benzenoid rings. The stability of the PAHs primarily depends on the arrangement of
the aromatic ring; the linear arrangement of aromatic rings (LMW-PAHs) represents the instability
and exhibits fewer recalcitrance characteristics, whereas the angular arrangement of aromatic rings
(HMW-PAHs) is highly stable and exhibits more recalcitrance features of PAHs compounds (Blumer
1976). In addition to this, according to United States Environmental Protection Agency (USEPA),
16 PAHs are classified as priority environmental pollutants and seven PAHs are categorized as
potent human carcinogens, known as carcinogenic PAHs (USEPA 2002).
PAHs are omnipresent environmental contaminants, generated from the partial burning of fossil
fuels. Air mass movement and transboundary deposition play a significant role in the distribution of
PAH in the environment. Through long-range transport, rural areas are also getting affected, which
are usually situated far from the actual origin of the PAHs. Soil and street dust are the primary sinks
for the deposition of atmospheric PAHs. Furthermore, this opens the gateway for PAHs to enter
the aquatic ecosystem. PAHs are preferentially fragmented and assembled in the particle state of
sediments in aquatic environments due to their hydrophobic nature. In this way, PAHs are present
throughout the multi-compartment structure of the ecosystem, paving the path for various exposure
routes to these carcinogens (Hussain et al. 2018). Broadly, PAHs can evolve from two types of
sources, i.e., naturogenic or anthropogenic sources. The naturogenic sources include forest fires,
volcanic eruptions, petroleum spills, bacterial and algal synthesis and decomposition of litter fall
(Abdel-Shafy and Mansour 2016). Various sources of PAHs and their exposure routes to human are
reported by many researchers. The major contributors of PAHs in the ecosphere are anthropogenic
sources such as the combustion of wood gas, fuels, crude oil and industrial wastes (Hussain
et al. 2018). PAHs produced from anthropogenic sources are mainly categorized into pyrogenic,
petrogenic and biological. Pyrogenic mediated PAHs are produced from partial burning of organic
matter subjected to high-temperature ranges from 350 to 1200°C under anaerobic conditions.
Examples of pyrogenic processes include the partial combustion of motors fuels in automobiles
and coal is thermal distilled into coal tar and coke, thermal cracking of oil deposits asphalt creation.
Therefore, pyrogenic PAHs released in the open air are accumulated more in urban regions. The
PAHs generated from the petrogenic process are similar to pyrogenic except regarding petroleum
processing (Nayak et al. 2022).
In this case, bioremediation is nature’s own green machinery that acts consistently as a good
cleanup, cost-effective, energy-efficient and eco-sustainable alternative equipment in comparison to
the physicochemical techniques viz, soil replacement, soil washing and flushing, chemical reduction
and oxidation, incineration and thermal desorption, vitrification, encapsulation, immobilization,
electrokinetic remediation, nonthermal plasma technology, etc. (Mandree et al. 2021, Kuppusamy
et al. 2017). The main objective of the bioremediation approach is based on the mineralization
of these hazardous compounds into non-toxic compounds and can be achieved by employing
bioremediating representatives like plants (phytoremediation), earthworms (vermiremediation),
as well as microorganisms, i.e., bacteria, yeast, fungi and algae. Such microbial agents are well
equipped with their enzymatic systems to bio transform the PAHs compounds into either carbon
dioxide (CO2) or partially degraded non-toxic metabolites, i.e., byproduct where no CO2 is liberated
(Cerniglia 1993). The microbial PAHs metabolism is generally achieved by a variety of mechanisms
based on the enzymatic depository systems.
88 Bioremediation for Sustainable Environmental Cleanup
Figure 6.1. Keywords co-occurrence network analysis using VOS viewer software.
as nature’s own reclamation or biorestoration practice where such pollutants are either degraded
or transformed into less toxic composites by various living organisms. Usually microorganisms
play an indispensable role in altering the pollutants into CO2, water, and cell biomass (Langenbach
2013). Examples of such bioremediation processes are usually known by different formulations
able. 6.1. Structure of PAHs, their molecular weight, solubility and health risk (Adopted from
Table. 6.1. Structure of PAHs, their molecular weight, solubility and health risk (Adopted from
Pathak et al. 2022, Hussain et al. 2018)
Pathak et al. 2022, Hussain et al. 2018)
90 Bioremediation for Sustainable Environmental Cleanup
PAHs Table.
Table.
Table.
Table. No.
6.1.
6.1.
6.1. ofStructure
StructureMolecular
Structure of
ofPAHs,PAHs,
PAHs, Solubility
their
their molecular
molecular weight,
weight, Health
solubility risk
solubility and and health
health risk
risk Structure
(Adopted
(Adopted from
from
PAHsTable. 6.1.
6.1. Structure
No. of of
of
of Molecular
Structure PAHs,
PAHs, their
their
their molecular
molecular
Solubility
molecular weight,
weight,
weight, solubility
solubility
Health
solubility and
risk
and and health
health
health risk
risk
risk (Adopted
(Adopted
Structure
(Adopted from
from
from
Table.6.1.
Table Rings
6.1. Pathak
6.1.Pathak
Structure
Structure
Pathak
Pathak ofweight
et
of al.
PAHs,
etal.
al. 2022,
PAHs,
2022,their their (mg/L)
Hussain
molecular
Hussain et
al.al.
molecular
weight,
etal. 2018)
weight,solubility
solubility
2018) solubilityand
and health riskandhealthhealthrisk
(Adopted risk(Adopted
from (Adoptedfrom
Pathak et al. fromHussain
2022,
Table. Rings
Structure
Pathak
Table. 6.1.Pathak
Structureet et
et of
al. al.
of
2022,
2022,
weight
PAHs,
2022,
PAHs,
Hussain
Hussain
their
Hussain
their et et
et al.al.
(mg/L)
molecular
molecular
2018)
2018)
weight,
2018) weight, solubility and health risk (Adopted from
(g/mol) et al. 2018).
Pathak
PAHs
Pathak etet
et
al.
al.
al.
2022,Hussain
(g/mol)
2022,
No.
2022,of
Hussainetetal.
Molecular
Hussain et
al.2018)
al.
2018)
Solubility
2018) Health risk Structure
PAHs
PAHs
PAHs No.
No.No. of
of of Molecular
Molecular
Molecular Solubility
Liver andHealth
Solubility
Solubility Health
kidney
Health
Health risk
failure,
riskrisk Structure
Structure
Structure
PAHs
PAHs No.of
No. of Molecular
Rings Molecular Solubility
weight Solubility (mg/L)Liver andHealth Healthriskrisk
kidney failure, Structure
Structure
Naphthalene PAHsPAHs2 Rings
No.
Ringsof Molecular
Rings
128.17 weight
Molecular
weight
weight 3.93 (mg/L)
Solubility
(mg/L)
(mg/L) anemia and skin risk Structure
No. of
Rings
Rings weightweight
(g/mol) Solubility
(mg/L)
(mg/L) Health risk Structure
Naphthalene PAHs 2 No.
Rings 128.17
of (g/mol)
Molecular
(g/mol)
weight 3.93
Solubility
(g/mol) (mg/L)
(g/mol) (mg/L) anemia and skin
Health risk Structure
Naphthalene Rings Rings
2 weight
(g/mol)
weight
128.17
(g/mol) 3.93 inflammation
(mg/L) Liver
Liver
inflammation
Liver
Liver and
and
and
and kidney
kidney
kidney
kidney failure,
failure,
failure,
(g/mol) Liver
Liver and
and kidney
kidney failure,
failure,
Naphthalene
Naphthalene 2 22 (g/mol)
128.17
128.17 3.93Carcinogenicity,
3.93 failure,
anemia
Liver anemia
anemia
and and and infant
and
skin
kidney skin
skinfailure,
Naphthalene
Naphthalene
Naphthalene
Acenaphthene 3
2
2
154.21
128.17
128.17
128.17 3.93 Carcinogenicity,
3.93
3.93 Liveranemia
anemia
anemia and and
andand skininfant
skin
kidney skinfailure,
128.17 1.93 3.93asthma and infection in
Liver inflammation
inflammation
and kidney
Naphthalene
Acenaphthene
Naphthalene 3 2 2 154.21 128.17 1.93 3.93 asthma inflammation
anemia
anemia andand
inflammation
inflammation
and skin failure,
infection
skin in
inflammation
3.93liver inflammation
Naphthalene
Acenaphthene 32 128.17
154.21 1.93 and
anemia kidney
and
Carcinogenicity,
Carcinogenicity, skin
Carcinogenicity,
Carcinogenicity,
Carcinogenicity, infant infant
infant
infant
infant
Acenaphthene liver and
inflammation kidney
Carcinogenicity,
inflammation infant
Acenaphthene
Acenaphthene 3 3 154.21
154.21 1.93
1.93Animalasthma asthma
asthma and
carcinogenicity
and and
infection
and
Carcinogenicity, infection
ininfant
infection in in
liverin
Acenaphthene
Acenaphthene 33
3 154.21
154.21
154.21 1.93
1.93
1.93 Animal
asthma
asthma
Carcinogenicity,
asthma andand infection
infection
infant
infection
carcinogenicity
liver and
Carcinogenicity,
and kidney kidney infant in in
Acenaphthene
Acenaphthylene 3 3 154.21 1.93 liver
asthma
liver
liver and
and
andand kidney
infection
kidney
kidney in
Acenaphthene
Acenaphthylene
Acenaphthene3 33152.1
152.1154.21
3.93
154.21 3.931.93 1.93 asthma
liver
asthma andkidney
and infection
infectioninin
liver
Animalandand
Animal
Animal
Animal carcinogenicity
kidney
carcinogenicity
carcinogenicity
carcinogenicity
3.93 liver and kidney
Acenaphthylene 3 152.1 3.93 Animal
Animal carcinogenicity
carcinogenicity
Acenaphthylene
Acenaphthylene
Acenaphthylene 3 3 152.1
152.1 3.93 liver
Animal andcarcinogenicity
kidney
Acenaphthylene
Acenaphthylene 33
3 152.1
152.1
152.1 3.93
3.93
3.93 Animal carcinogenicity
Animal carcinogenicity
Acenaphthylene
Acenaphthylene 3
33 152.1
152.1 3.93
3.93 Diarrhea, nausea, eye
Acenaphthylene 152.1 3.93 Diarrhea, nausea, eye
Anthracene 3 178.23 0.076 infection and nausea,
Diarrhea,
Diarrhea,
Diarrhea,
Diarrhea, skinnausea,
nausea,
nausea, eyeeye
eye
Anthracene Anthracene 3 178.23 0.076 0.076 Diarrhea,
infection
Diarrhea, andnausea,
skin
nausea, eyeeye
eye
Anthracene
Anthracene
Anthracene
Anthracene 333 3 3 178.23
178.23
178.23
178.23
178.23 0.076
0.076 irritation
0.076irritation
0.076 infection
Diarrhea,
infection
infection
infection
infection and
and and and
nausea,
and
skin skinskin
eye
skin
skin
irritation
Anthracene 3 178.23 0.076 Diarrhea,
infection nausea,
irritation
Diarrhea, and
nausea,skineye
eye
Anthracene
Anthracene 333 178.23
178.23 0.076
0.076 Carcinogenicity
irritation
infection
irritation
irritation
infection
irritation
and
and and
skin
skin skin
Anthracene 178.23 0.076 Carcinogenicity
infection
irritation
Carcinogenicity
Carcinogenicity
and
and skinand
Carcinogenicity andskin
and
and skin
skin
skin
Phenanthrene 3
Phenanthrene 3178.23 178.23 1.2 1.2 irritation Carcinogenicity
irritation
Carcinogenicity
Carcinogenicity and and
skin skin
skin
Phenanthrene
Phenanthrene
Phenanthrene
Phenanthrene 3
Phenanthrene 3 178.23
33 3 178.23
178.23
178.23
178.23 1.2 1.2
1.2 1.2
1.2 irritation
irritation
irritation
irritation
Carcinogenicity
irritation
irritation
irritation and skin
Phenanthrene 3 178.23 1.2 Carcinogenicity
irritation
Carcinogenicity and
andskinskin
Phenanthrene 3 178.23 1.2 Changes irritation
in morphology
Phenanthrene
Phenanthrene 33 178.23
178.23 1.2
1.2 Changesirritation
Changes
Changes
irritation
Changesin in morphology
morphology
inin
morphology
morphology
Changes
Changes in morphology
in morphology
Pyrene Pyrene 4
Pyrene 4202.25
4 44 0.0770.077
202.25
202.25 of blood
0.077 ofcell
blood cell of
Pyrene
Pyrene Pyrene
Pyrene 4 4 202.25
4 202.25
202.25
202.25 0.077
0.0770.077
0.077 ofChanges
ofof
Changes
blood
of blood
Changes
in morphology
blood in
cell
blood
in
cell
cellmorphology
cell
Pyrene
Pyrene 4
202.25
202.25
0.077
0.077
of
of
blood
Changes
bloodbloodinmorphology
cell
cell morphology
cell
Pyrene
Pyrene 44 202.25
202.25 0.077
0.077 ofofblood
bloodcellcell
Fluorene Fluorene
Fluorene
Fluorene
Fluorene 3
Fluorene 33 33
3166.22 166.22
166.22
166.22
166.22
166.22 1.68–1.98
1.68–1.98
1.68–1.981.68–1.98
1.68–1.98
1.68–1.98 Eye
Eye
Eye andEye
Eye
Eye and
skin
and
and and
and skin
skin
infection
skin
skin
skin infection
infection
infection
infection
infection
Fluorene Fluorene 3 3 166.22166.221.68–1.981.68–1.98 Eye and skin infection
Fluorene
Fluorene 333 166.22
166.22 1.68–1.98Eye
1.68–1.98 Eyeand and skin
skininfection
infection
Fluorene 166.22 1.68–1.98 Eye Eyeand andskinskininfection
infection
Fluoranthene
Fluoranthene
Fluoranthene
Fluoranthene 44 44 202.25 0.2–2.6
202.25
202.25
202.25 0.2–2.6 Acute
0.2–2.6
0.2–2.6 Acute
Acute
Acute animal
animal
animal
animal toxicity
toxicity
toxicity
toxicity
Fluoranthene
Fluoranthene
FluorantheneFluoranthene 4 4202.25
4 202.25
202.250.2–2.6 0.2–2.6
0.2–2.6 Acute Acute
Acute
animal animal
animal toxicity
toxicitytoxicity
Fluoranthene
Fluoranthene
Fluoranthene
4 444 202.25 202.25 0.2–2.6
202.25
202.25
0.2–2.6 Acute
0.2–2.6
0.2–2.6 Acute
Acute animaltoxicity
animal
Acuteanimal
animaltoxicity
toxicity
toxicity
Benzo[a]
Benzo[a]
Benzo[a]
Benzo[a] Animal
Animal
Animal
Animal carcinogenicity
carcinogenicity
carcinogenicity
carcinogenicity
Benzo[a]
anthracene Animal carcinogenicity
Benzo[a] Benzo[a]
anthracene
Benzo[a]
anthracene 4 4 228.3
228.3 0.01
0.01Animal andand
Animal skinskin
carcinogenicityirritation
irritation
carcinogenicity
anthracene
Benzo[a]
anthracene
Benzo[a] anthracene
Benzo[a]
4444 228.3
228.3
228.3
228.3 0.01
0.01
0.01
0.01 and
Animal
Animal
and
Animal
and
Animal
skin
skin irritation
irritation
carcinogenicity
skincarcinogenicity
irritation
carcinogenicity
carcinogenicity
and
anthracene anthracene 4
anthracene 4228.3 228.3 0.01 0.01and and
skin skin irritation
irritation
skin irritation
anthraceneanthracene 4 44 228.3228.3
228.3 0.010.01
0.01 and and
and skin
skinskin irritation
Carcinogenicity,
irritation
Carcinogenicity,
irritationemesis,
Carcinogenicity, emesis,
emesis,
emesis,
Carcinogenicity,
Carcinogenicity, emesis,
Chrysene
Chrysene
Chrysene
Chrysene diarrhea,
Carcinogenicity,
diarrhea,
diarrhea,
diarrhea, eye
eye eye
eye and
and and
and skin
emesis,
skin
skin
skin
Chrysene 444 44 228.29 0.0028 Carcinogenicity,
0.0028 Carcinogenicity,
diarrhea,
infection,
Carcinogenicity, eye emesis,
andemesis,
heart skin
problems
emesis,
ChryseneChrysene 228.29
228.29
228.29
228.29 0.0028
0.0028
0.0028 Carcinogenicity,
Carcinogenicity,
infection,
diarrhea,
infection,
infection, eyeheart
and
heart
heart emesis,
emesis,
problems
skin
problems
problems
Chrysene Chrysene 4 228.29 0.0028 diarrhea,
0.0028diarrhea, eyeeye
infection, and
heart
and skin
problems
skin
Chrysene Chrysene 4
4228.29 228.29
228.29 0.0028
diarrhea,
diarrhea,
infection,
diarrhea,
infection, eye eye
eye
heart
heart
and
and
and skin
skin
problems
skin
problems
Benzo[b]
Benzo[b]
Benzo[b]
Benzo[b] 4 4 228.290.00280.0028 infection,
infection,
infection,heart
heart problems
heart
Carcinogenicity,
Carcinogenicity,
Carcinogenicity,
Carcinogenicity,
problems
problems
Benzo[b] 4 228.29 0.0028 infection, heart problems
Carcinogenicity,
fluoranthene
fluoranthene
Benzo[b]
fluoranthene
fluoranthene 55 55 252.3
252.3
252.3
252.3 0.0012 Carcinogenicity,
0.0012
0.0012
0.0012 jaundice,
jaundice,
jaundice,
jaundice, liverliver
liver
liver and
and and
and
Benzo[b]
fluoranthene
Benzo[b] 5 252.3 0.0012 Carcinogenicity,
jaundice,
kidney
Carcinogenicity, liver
failure and
Benzo[b] fluoranthene
fluoranthene
Benzo[b] 55
55 252.3
252.3
252.3 0.0012
0.0012 Carcinogenicity,
kidney
jaundice,
Carcinogenicity,
kidney
kidney
0.0012 Carcinogenicity,
jaundice,
kidney
failure
liver
failure
failure
liver
failure
and
jaundice,
and
Benzo[b] Benz[k]
luoranthenefluoranthene
Benz[k]
fluoranthene
Benz[k]
Benz[k]
Benz[k] 5 252.3 252.3 0.00120.0012jaundice, jaundice,
liver
kidney liver
Carcinogenicity,
andfailure
kidney
Carcinogenicity,
liver
Carcinogenicity,
Carcinogenicity,
kidney failure
Carcinogenicity,
andliver
andfailure liver
liver
liver
liver
fluoranthene fluoranthene
fluoranthene
Benz[k]
fluoranthene
fluoranthene 5 555252.3
5 252.3
252.3 0.0012
252.3
252.3 0.00076
0.00076
0.00076 jaundice,
0.00076 kidney
and
and and
and liver
failure
kidney
kidney
Carcinogenicity,
kidney
kidney andfailure
failure liver
failure
failure
Benz[k]
fluoranthene 5 252.3 0.00076 kidney andfailure
Carcinogenicity,
kidney failureliver
Benz[k]
fluoranthene 5 252.3 0.00076 kidney and failurefailure
Carcinogenicity,
kidney liver
Benz[k] fluoranthene 555 252.3 0.00076 Carcinogenicity,
and kidney failureliver
Benz[k] fluoranthene
Benz[k] 252.3
252.3 0.00076 and kidney failureand
0.00076 Carcinogenicity,
Carcinogenicity, liver liver
luoranthene 5
fluoranthene 252.3 0.00076 and kidney failure
kidney failure
fluoranthene 5 252.3 0.00076 and kidney failure
Carcinogenicity,
Carcinogenicity,
Carcinogenicity, nausea,
nausea,
nausea,
Indeno[1,2,3-cd] 6 276.3 0.062 Carcinogenicity, nausea,
2,3-
2,3-
[1,2,3- 66 6 pyrene 276.3
276.3
276.3 0.062
0.062
0.062 emesis
emesis and
and diarrhea
diarrhea
emesis and diarrhea
emesis and diarrhea
eeene
viz, biosorption used dead or alive microbial organic matter, bioaccumulation assisted by microbes,
biostimulation is ameliorating on-site microbial community, phytoremediation associated with plants,
bioaugmentation is the unnatural incorporation of microbial communities and rhizoremediation is
the interaction between plants and microbes. Among the bioremediation practices, microbes (algae,
fungi, bacteria) mediated remediation is the most effective in PAHs mineralization and numerous
documentations have reported more than 100 genera and 200 species of microorganisms. Where,
fungi are the most predominant group with 103 genera among the other microbial representatives,
i.e., bacteria (79 genera), algae (19 genera) and cyanobacteria (9 genera) documented from a
different environment and responsible for effective PAHs degradation (Premnath et al. 2021).
Fungi and bacteria utilize versatile strategies for remediating PAHs in contaminated soil. The first
stage of the bacterial PAHs degradation mechanism includes oxidation as well as hydroxylation of
PAHs. The detoxification operation is carried out as the initial stage in fungal PAHs mineralization.
Individually, fungal members like Aspergillus sp., Fusarium oxysporum and Trichocladium
canadense can also mineralize LMW-PAHs. However, in contrast to the LMW-PAHs, HMW-PAHs,
i.e., pyrene (PYR) and phenanthrene (PHE) can be efficiently mineralized by the Penicillium sp.,
A. terreus and Verticillium sp. (Biswas et al. 2015). Moreover, as a consortium of bacteria and
fungi, rather than individuals, proficiently enhance the rate of PAH degradation (Winquist et al.
2014, Li and Li 2008). On the other hand, fungal consortia of A. flavus, A. fumigatus, A. nomius,
Trichoderma asperellum and Rhizomucor variabilis can also escalate the remediation of Benzo(a)
pyrene [B(a)p], PYR and PHE (Tripathi et al. 2017). Universally fungal agents are equipped with
several remediating strategies like enzymatic activities, the production of biosurfactants and the
utilization of rhizospheric and endophytic fungal communities.
and Kumar 2019). Ligninolytic fungi are proficient in synthesizing extracellular enzymes viz, LiP
and MnP which are utilized in PAHs mineralization, apart from intracellular enzymes. Ligninolytic
enzymes, break down such contaminants and produce free radicals. These free radicals then oxidize
various aromatic molecules, resulting in different quinones (Kadri et al. 2017). Non-ligninolytic
fungi are usually inefficient in extracellular enzyme production and intracellular enzymes such
as CYP450 play a significant role in PAHs degradation (Dutta and Laha 2022). The predominant
fungal phyla that are highly considered as a potential bioagent for PAHs remediation belong to
Ascomycota (non-ligninolytic fungi), followed by Basidiomycota (ligninolytic fungi) and the least
by Zygomycota that is categorized under the non-ligninolytic members (Banerjee and Mandal 2020,
Passarini et al. 2011). In this, diversified fungal agents which have the potential to eliminate the
PAHs from contaminated soil and their mode of action have been documented in (Table 6.2) and
from this analysis a similar pattern of dominant fungal phyla was observed, where Ascomycota
appeared as the most dominant member (64%) followed by Basidiomycota (33%) and Zygomycota
(3%), respectively (Figure 6.3).
white-rot fungus can degrade LMW-PAHs as well as HMW-PAHs with removal efficiencies of up
to 58–73 and 21–26%, respectively (Kariyawasam et al. 2021, Leonardi et al. 2007). Utilization
of extracellular enzymes laccases and peroxidases accelerate the B(a)P metabolized into B(a)
P-1,6, -3,6-quinones or B(a)P-6,12-quinones intermediates observed in fungal species such as
P. chrysosporium, Cunninghamella elegans, A. ochraceus, T. versicolor and P. cinnabarinus
illustrated in Figure 6.4 (Majcherczyk et al. 1998, Datta and Samanta 1988, Haemmerli et al. 1986,
Cerniglia and Gibson 1980). Another report of WRF P. ostreatus in immobilized commercial pellets
that can degrade a significant number of PAHs such as 69.1% of Benzo(a)anthracene [B(a)A], 29.7%
of chrysene (CHY), 39.7% of Benzo(b)fluoranthene [B(b)F], 32.8% of Benzo (k)fluoranthene
[B(k)F], 85.2% of B(a)P and 80% of the total PAHs (Covino et al. 2010). Furthermore, Bhattacharya
et al. (2017) discovered that the biodegradation of B(a)P was enhanced when it was subjected to
96 Bioremediation for Sustainable Environmental Cleanup
Figure 6.4. B(a)P mineralization by P. chrysosporium, C. elegans, A. ochraceus, T. versicolor and P. cinnabarinus
(Majcherczyk et al. 1998, Datta and Samanta 1988, Haemmerli et al. 1986, Cerniglia and Gibson 1980).
consortia of P. ostreatus with Penicillium chrysogenum by 86.15% removal rate and P. ostreatus with
Pseudomonas aeruginosa by 75.1% removal rate, whereas P. ostreatus PO-3 only remediated 64.3%.
On the other hand, the ligninolytic fungi or the WRF are known for their magnificent PAHs
degrading abilities with their potential intra and extracellular enzymes. In WRF mediated PAHs
remediation process, PAHs oxidation begins with the formation of dihydrodiol by a multi-component
dioxygenase enzyme system which further undergoes ortho or meta ring cleavage pathways and
produces metabolites such as catechol and protocatechuate (Sipilä et al. 2010). Furthermore, some
fungi undergo more extensive PAH decomposition, leading to the breakdown of the benzenoid rings
and the release of CO2. Enzymes involved in WRF-mediated PAHs breakdown mainly belong to the
LiP, laccase, MnP and VP (Gupta and Pathak 2020). The mechanisms involved in PAHs degradation
by both ligninolytic and non-ligninolytic fungi have been depicted in Figure 6.5.
6.3.1.1.1 Laccase
Extracellular enzymes, especially laccase, are crucial features that contribute to PAHs cleanup.
Laccases are benzenediol: oxygen oxidoreductases and are metalloproteins in nature, which belong
to the polyphenol oxidases and consist of four copper atoms of altered types in their catalytic site
(Baldrian 2006). Laccases are one of the key ligninolytic biomolecules that oxidize a broad range of
aromatic compounds holding phenolic moieties, benzenothiols, aromatic amines and hydroxylindols
(Dhagat and Jujjavarapu 2022). Such enzymes utilize the molecular oxygen as an electron acceptor
and organic or inorganic metal complexes as substrates (Zimmerman et al. 2008). Additionally,
laccases use oxygen as an oxidizing agent and reduce it into water molecules (Tavares et al.
2006). Laccases are produced by diversified living systems like plants, fungi, bacteria and insects
(Show et al. 2022). Among microbes, laccases of bacterial origin are reported from some potent
Fungal Strategies for the Remediation of Polycyclic Aromatic Hydrocarbons 97
Figure 6.5. Ligninolytic and non-ligninolytic fungal mechanism responsible for PAHs degradation.
members viz, Bacillus subtilis, Azospirillum lipoferum, Sinorhizobium meliloti and Streptomyces
lavendulae, whereas in the fungal system, intracellular and extracellular laccases are documented for
a different type of PAHs remediation (Banerjee and Mandal 2020, Pawlik et al. 2016). Several reports
have demonstrated that fungi utilize PAHs via the production of an extracellular enzyme like laccases
resulting in PAHs degradation with CO2 as a byproduct. A significant amount of PAHs removal
capability by laccase-producing WRF Anthracophyllum discolor was documented with 62% of PHE,
73% of anthracene (ANTH), 54% of fluoranthene (FLU), 60% of PYR and 75% of B(a)P degrading
potentiality (Acevedo et al. 2011). In a recent study, a WRF strain Trametes versicolor exhibited
excellent remediational attributes like 81.0% CHY and 91.0% B(a)P degradation with
37.8 Ug–1 laccase producing potentiality (Vipotnik et al. 2022). In addition, another report of a WRF
Pycnoporus sanguineus with 2516.7 UL–1 of laccase production ability displayed 90.1% B(a)A and
45.6% PHE degrading capacity (Li et al. 2018). Further, laccase produced by T. versicolor was also
found to be able to remediate contaminants such as acenaphthene (ACE), B(b)F and PYR (Noman
et al. 2019). In this regard, Punnapayak et al. (2009) reported that laccase-producing WRF
Ganoderma lucidum can also efficiently degrade ACE and acenaphthylene.
6.3.1.1.2 Peroxidases
Peroxidases are the glycosylated extracellular enzyme that requires hydrogen peroxide to catalyze
lignin and organopollutant compounds like PAHs (Thurston 1994). The peroxidases are classified
into two categories based on their substrate interaction—MnP which is the most effective reducing
substrate and LiP which catalyzes both non-aromatic and aromatic compounds (Ten Have and
Teunissen 2001). Hydrogen peroxide (H2O2) is essential for the activation of the MnP and LiP are
synthesized by fungal producers during their metabolic processes. In such metabolic processes,
H2O2 acts as an oxidizing agent that oxidizes the PAHs and transforms them into a substrate
98 Bioremediation for Sustainable Environmental Cleanup
(Deshmukh et al. 2016). VPs also fall in this peroxidase isozymes family, capable to operate on
a number of substrates. Several scientific reports have been documented regarding the PAHs
degradation by fungal peroxidases.
soluble and non-hazardous when compared with their parental PAHs compounds, hence lowering
health-risk. C. elegans exhibit a sulfur addition process in which naphthalene (NAP) and ANTH are
converted into 1-naphthyl and 1-anthryl sulfate metabolites, respectively (Lisowska and Długoński
2003). On the other hand, enhancement in PAHs degradation by accelerating the PAHs solubility
can be achieved by conjugation process. P. chrysosporium shows glucose as PAHs conjugates.
According to Sutherland (1992), 7,12-dimethylbenzanthracene can be biotransformed by C. elegans
producing trans-3,4-dihydrodiol metabolites. It can be further metabolized into phenol carried out
by Mucor sp., P. chrysogenum. S. racemosum biosystem via conjugating with glucuronic acid.
ANTH biodegradation can be accomplished by various fungi such as Fusarium sp., Aspergillus sp.,
Penicillium sp., Trichoderma sp., Ulocladium chartarum and Absidia cylindrospora (Giraud et al.
2001). In addition to this, the non-ligninolytic fungi involved in pherenthrene metabolism usually
belong to Alternaria sp., F. culmorum, P. janczewskii, C. elegans, Cladosporium herbarum and
T. hamatum, etc. (Schmidt et al. 2010).
plants are omnipresent. Arbuscular mycorrhizae perform several roles in plant growth, such as
enhancing nutrient and minerals uptake, drought resistance and immunization from plant diseases
(Pramanik et al. 2021, Hildebrandt et al. 2007). Arbuscular mycorrhizae such as Glomus mosseae
and G. etunicatum associated with ryegrass roots efficiently degrade PAHs such as FLR and PHE
(Gao et al. 2010). Chulalaksananukul et al. (2006) reported that Fusarium sp., isolated from the
leaves of Pterocarpus macrocarpus, capable of degrading B(a)P by 70% of 100 ppm concentration.
Kannangara et al. (2016) identified different endophytic fungi such as P. oxalicum, Nigrospora
oryzae, A. oryzae and A. aculeatus which are very efficient at degrading NAP as well as PHE above
than 85% of removal efficiency.
6.7 Conclusion
Nowadays, bioremediation approaches for environmental pollutants like PAHs of contaminated soil
have received significant importance in the scientific community as they are not easily eliminated
by natural processes and can reside for decades in the environment. In accordance with this, PAHs
are associated with various health issues such as cancer, kidney failure, lung malfunction, etc. In
this context, among the different PAHs remediation approaches, fungal bioremediation approaches
are the most promising techniques for the future as it promotes sustainability and cost-effectivity.
Fungal representatives like Basidiomycota and Ascomycota are the predominant communities,
involved in biodegradation of different types of LMW and HMW PAHs compounds. Accelerated
biotransformation of PAHs via mycoremediation strategies such as enzymatic metabolism,
biosurfactant production, biochar immobilized fungal administration and utilization of rhizospheric
and endophytic fungi can be harnessed as nature’s own medicine to combat with such polluting
ecological malaises. Utilization of these fungal strategies can accelerate the bioremediation of PAHs
and can open a “novel biotechnological green-window” for eco-sustainable mother earth.
Acknowledgement
Sandipan Banerjee and Nitu Gupta thank the Department of Biotechnology, Govt. of India, for
granting DBT Twinning Project and Research Fellowship [No. BT/PR25738/NER/95/1329/2017
dated December 24, 2018]. Apurba Koley is thankful to the BBSRC, United Kingdom, for granting
funding from the BEFWAM project: Bioenergy, Fertilizer and Clean water from Invasive Aquatic
macrophytes [Grant Ref: BB/S011439/1] for financial support and research fellowship.
References
Abdel-Shafy, H. I. and M. S. Mansour. 2016. A review on polycyclic aromatic hydrocarbons: source, environmental
impact, effect on human health and remediation. Egypt. J. Pet. 25(1): 107–123.
Acevedo, F., L. Pizzul, M. del Pilar Castillo, R. Cuevas and M. C. Diez. 2011. Degradation of polycyclic aromatic
hydrocarbons by the Chilean white-rot fungus Anthracophyllum discolor. J. Hazard. Mater. 185(1): 212–219.
Agrawal, N., P. Verma and S. K. Shahi. 2018. Degradation of polycyclic aromatic hydrocarbons (phenanthrene and
pyrene) by the ligninolytic fungi Ganoderma lucidum isolated from the hardwood
stump. Bioresour. Bioprocess. 5(1): 1–9.
Anderson, T. A., E. A. Guthrie and B. T. Walton. 1993. Bioremediation in the rhizosphere. Environ. Sci. Technol. 27(13):
2630–2636.
Aranda, E., P. Godoy, R. Reina, M. Badia-Fabregat, M. Rosell, E. Marco-Urrea and I. García-Romera. 2017. Isolation
of Ascomycota fungi with capability to transform PAHs: Insights into the biodegradation mechanisms of
Penicillium oxalicum. Int. Biodeterior. Biodegrad. 122: 141–150.
Arun, A., P. P. Raja, R. Arthi, M. Ananthi, K. S. Kumar and M. Eyini. 2008. Polycyclic aromatic hydrocarbons (PAHs)
biodegradation by basidiomycetes fungi, Pseudomonas isolate, and their cocultures: comparative in vivo and
in silico approach. Appl. Biochem. Biotechnol. 151(2): 132–142.
Arun, A. and M. Eyini. 2011. Comparative studies on lignin and polycyclic aromatic hydrocarbons degradation by
basidiomycetes fungi. Bioresour. Technol. 102(17): 8063–8070.
Fungal Strategies for the Remediation of Polycyclic Aromatic Hydrocarbons 103
Aydin, S., H. A. Karaçay, A. Shahi, S. Gökçe, B. Ince and O. Ince. 2017. Aerobic and anaerobic fungal metabolism
and Omics insights for increasing polycyclic aromatic hydrocarbons biodegradation. Fungal Biol. Rev. 31(2):
61–72.
Baldrian, P. 2006. Fungal laccases–occurrence and properties. FEMS Microbiol. Rev. 30(2): 215–242.
Ball, A. and A. Truskewycz. 2013. Polyaromatic hydrocarbon exposure: an ecological impact
ambiguity. Environ. Sci. Pollut. Res. 20(7): 4311–4326.
Banerjee, S., T. K. Maiti and R. N. Roy. 2017. Protease production by thermo-alkaliphilic novel gut isolate Kitasatospora
cheerisanensis GAP 12.4 from Gryllotalpa africana. Biocatal. Biotransformation. 35(3): 168–176.
Banerjee, S. and N. C. Mandal. 2020. Fungal Bioagents in the remediation of degraded soils. In Microbial Services in
Restoration Ecology. pp. 191–205. Elsevier.
Bernhardt, R. 2006. Cytochromes P450 as versatile biocatalysts. J. Biotechnol. 124(1): 128–145.
Bezalel, L., Y. Hadar, P. P. Fu, J. P. Freeman and C. E. Cerniglia. 1996. Initial oxidation products in the metabolism
of pyrene, anthracene, fluorene, and dibenzothiophene by the white rot fungus Pleurotus ostreatus. Appl.
Environ. Microbiol. 62: 2554–2559.
Bhattacharya, S. S., K. Syed, J. Shann and J. S. Yadav. 2013. A novel P450-initiated biphasic process for sustainable
biodegradation of benzo [a] pyrene in soil under nutrient-sufficient conditions by the white rot fungus
Phanerochaete chrysosporium. J. Hazard. Mater. 261: 675–683.
Bhattacharya, S., A. Das, M. Palaniswamy and J. Angayarkanni. 2017. Degradation of benzo [a] pyrene by Pleurotus
ostreatus PO‐3 in the presence of defined fungal and bacterial co‐cultures. J. Basic Microbiol. 57(2): 95–103.
Biswas, B., B. Sarkar, R. Rusmin and R. Naidu. 2015. Bioremediation of PAHs and VOCs: advances in clay mineral–
microbial interaction. Environ. Int. 85: 168–181.
Blumer, M. 1976. Polycyclic aromatic compounds in nature. Sci. Am. 234: 34–45.
Bogan, B. W., R. T. Lamar and K. E. Hammel. 1996. Fluorene oxidation in vivo by Phanerochaete chrysosporium
and in vitro during manganese peroxidase-dependent lipid peroxidation. Appl. Environ. Microbiol. 62(5):
1788–1792.
Boonchan, S., M. L. Britz and G. A. Stanley. 2000. Degradation and mineralization of high-molecular-weight
polycyclic aromatic hydrocarbons by defined fungal-bacterial cocultures. Appl. Environ. Microbiol. 66(3):
1007–1019.
Cajthaml, T., P. Erbanová, V. Šašek and M. Moeder. 2006. Breakdown products on metabolic pathway of degradation
of benz [a] anthracene by a ligninolytic fungus. Chemosphere. 64(4): 560–564.
Cajthaml, T., P. Erbanová, A. Kollmann, Č. Novotný, V. Šašek and C. Mougin. 2008. Degradation of PAHs by
ligninolytic enzymes of Irpex lacteus. Folia Microbiol. 53(4): 289–294.
Cao, H., C. Wang, H. Liu, W. Jia and H. Sun. 2020. Enzyme activities during benzo(a) pyrene degradation by the
fungus Lasiodiplodia theobromae isolated from a polluted soil. Sci. Rep. 10: 865. https://doi.org/10.1038/
s41598-020-57692-6.
Capotorti, G., P. Cesti, A. Lombardi and G. Guglielmetti. 2005. Formation of sulfate conjugates metabolites in
the degradation of phenanthrene, anthracene, pyrene and benzo [a] pyrene by the ascomycete Aspergillus
terreus. Polycycl. Aromat. Comp. 25(3): 197–213.
Cerniglia, C. E. and D. T. Gibson. 1980. Oxidation of benzo [a] pyrene by the filamentous fungus Cunninghamella
elegans. J. Biol. Chem. 254(23): 12174–12180.
Cerniglia, C. E. 1993. Biodegradation of polycyclic aromatic hydrocarbons. Curr. Opin. Biotechnol. 4(3): 331–338.
Chowdhary, P., G. Shukla, G. Raj, L. F. R. Ferreira and R. N. Bharagava. 2019. Microbial manganese peroxidase: a
ligninolytic enzyme and its ample opportunities in research. SN Appl. Sci. 1(1): 1–12.
Choinowski, T., W. Blodig, K. H. Winterhalter and K. Piontek. 1999. The crystal structure of lignin peroxidase at 1.70
Å resolution reveals a hydroxy group on the Cβ of tryptophan 171: a novel radical site formed during the redox
cycle. J. Mol. Biol. 286(3): 809–827.
Chulalaksananukul, S., G. M. Gadd, P. Sangvanich, P. Sihanonth, J. Piapukiew and A. S. Vangnai. 2006. Biodegradation
of benzo (a) pyrene by a newly isolated Fusarium sp. FEMS Microbiol. Lett. 262(1): 99–106.
Collins, P. J. and A. D. Dobson. 1996. Oxidation of fluorene and phenanthrene by Mn (II) dependent peroxidase
activity in whole cultures of Trametes (Coriolus) versicolor. Biotechnol. Lett. 18(7): 801–804.
Conejo-Saucedo, U., D. R. Olicón-Hernández, T. Robledo-Mahón, H. P. Stein, C. Calvo and E. Aranda. 2019.
Bioremediation of polycyclic aromatic hydrocarbons (PAHs) contaminated soil through fungal communities.
In Recent Advancement in White Biotechnology Through Fungi. pp. 217–236. Springer, Cham.
Covino, S., K. Svobodová, M. Čvančarová, A. D’Annibale, M. Petruccioli, F. Federici et al. 2010. Inoculum carrier
and contaminant bioavailability affect fungal degradation performances of PAH-contaminated solid matrices
from a wood preservation plant. Chemosphere. 79(8): 855–864.
D’Annibale, A., M. Ricci, V. Leonardi, D. Quaratino, E. Mincione and M. Petruccioli. 2005. Degradation of aromatic
hydrocarbons by white-rot fungi in a historically contaminated soil. Biotechnol. Bioeng. 90: 723–731.
104 Bioremediation for Sustainable Environmental Cleanup
Dai, C. C., L. S. Tian, Y. T. Zhao, Y. Chen and H. Xie. 2010. Degradation of phenanthrene by the endophytic fungus
Ceratobasidum stevensii found in Bischofia polycarpa. Int. Biodeterior. 21(2): 245–255.
Datta, D. and T. B. Samanta. 1988. Effect of inducers on metabolism of benzo (a) pyrene in vivo and in vitro: analysis
by high pressure liquid chromatography. Biochem. Biophys. Res. Commun. 155(1): 493–502.
Dell’Anno, F., C. Sansone, A. Ianora and A. Dell’Anno. 2018. Biosurfactant-induced remediation of contaminated
marine sediments: current knowledge and future perspectives. Mar. Environ. Res. 137: 196–205.
Deshmukh, R., A. A. Khardenavis and H. J. Purohit. 2016. Diverse metabolic capacities of fungi for bioremediation.
Indian J. Microbiol. 56(3): 247–264.
Dhagat, S. and S. E. Jujjavarapu. 2022. Utility of lignin‐modifying enzymes: a green technology for organic compound
mycodegradation. J. Chem. Technol. Biotechnol. 97(2): 343–358.
Dutta, S. and S. Laha. 2022. An approach toward the biodegradation of PAHs by microbial consortia. In Development
in Wastewater Treatment Research and Processes. pp. 383–406. Elsevier.
Eibes, G., C. McCann, A. Pedezert, M. T. Moreira, G. Feijoo and J. M. Lema. 2010. Study of mass transfer and
biocatalyst stability for the enzymatic degradation of anthracene in a two-phase partitioning bioreactor.
Biochem. Eng. J. 51(1-2): 79–85.
Fayeulle, A., E. Veignie, R. Schroll, J. C. Munch and C. Rafin. 2019. PAH biodegradation by telluric saprotrophic
fungi isolated from aged PAH-contaminated soils in mineral medium and historically contaminated soil
microcosms. J. Soils Sediments. 19(7): 3056–3067.
Fenibo, E. O., G. N. Ijoma, R. Selvarajan and C. B. Chikere. 2019. Microbial surfactants: The next generation
multifunctional biomolecules for applications in the petroleum industry and its associated environmental
remediation. Microorganisms. 7(11): 581.
Field, J. A., E. deJong, G. F. Costa and J. A. M. deBont. 1992. Biodegradation of polycyclic aromatic hydrocarbons
by new iso-lates of white-rot fungi. Appl. Environ. Microbiol. 58: 2219–2226.
Gadd, G. M. and G. M. Gadd (Eds.). 2001. Fungi in bioremediation (No. 23). Cambridge University Press.
Gao, Y., Z. Cheng, W. Ling and J. Huang. 2010. Arbuscular mycorrhizal fungal hyphae contribute to the uptake of
polycyclic aromatic hydrocarbons by plant roots. Bioresour Technol. 101(18): 6895–6901.
García-Delgado, C., I. Alfaro-Barta and E. Eymar. 2015. Combination of biochar amendment and mycoremediation
for polycyclic aromatic hydrocarbons immobilization and biodegradation in creosote-contaminated soil. J.
Hazard. Mater. 285: 259–266.
Garon, D., S. Krivobok, D. Wouessidjewe and F. Seigle-Murandi. 2002. Influence of surfactants on solubilization and
fungal degradation of fluorene. Chemosphere. 47(3): 303–309.
Garon, D., L. Sage and F. Seigle-Murandi. 2004. Effects of fungal bioaugmentation and cyclodextrin amendment on
fluorene degradation in soil slurry. Biodegradation. 15(1): 1–8.
Ghosal, D., S. Ghosh, T. K. Dutta and Y. Ahn. 2016. Current state of knowledge in microbial degradation of polycyclic
aromatic hydrocarbons (PAHs): a review. Front Microbiol. 7: 1–27.
Giraud, F., P. Guiraud, M. Kadri, G. Blake and R. Steiman. 2001. Biodegradation of anthracene and fluoranthene
by fungi isolated from an experimental constructed wetland for wastewater treatment. Water Res. 35(17):
4126–4136.
Guo, J., X. Liu, X. Zhang, J. Wu, C. Chai, D. Ma et al. 2019. Immobilized lignin peroxidase on Fe3O4@ SiO2@
polydopamine nanoparticles for degradation of organic pollutants. Int. J. Biol. Macromol. 138: 433–440.
Guo, J. and X. Wen. 2021. Performance and kinetics of benzo (a) pyrene biodegradation in contaminated water and
soil and improvement of soil properties by biosurfactant amendment. Ecotoxicol. Environ. Saf. 207: 111292.
Gupta, S. and B. Pathak. 2020. Mycoremediation of polycyclic aromatic hydrocarbons. In Abatement of Environmental
Pollutants. pp. 127–149. Elsevier.
Haemmerli, S. D., M. S. Leisola, D. Sanglard and A. Fiechter. 1986. Oxidation of benzo (a) pyrene by extracellular
ligninases of Phanerochaete chrysosporium. Veratryl alcohol and stability of ligninase. J. Biol. Chem. 261(15):
6900–6903.
Henkel, M. and R. Hausmann. 2019. Diversity and classification of microbial surfactants. Biobased surfactants.
pp. 41–63. AOCS Press.
Hidayat, A. and S. Tachibana. 2015. Simple screening for potential chrysene degrading fungi. KnE Life Sci., 364–370.
Hildebrandt, U., M. Regvar and H. Bothe. 2007. Arbuscular mycorrhiza and heavy metal tolerance. Phytochemistry.
68: 139–146.
Hussain, K., R. R. Hoque, S. Balachandran, S. Medhi, M. G. Idris, M. Rahman and F. L. Hussain. 2018. Monitoring
and risk analysis of PAHs in the environment. Handbook of Environmental Materials Management, 1–35.
Imam, A., S. K. Suman, P. K. Kanaujia and A. Ray. 2022. Biological machinery for polycyclic aromatic hydrocarbons
degradation: A review. Bioresour. Technol. 343: 126121.
Jatav, H. S., V. D. Rajput, T. Minkina, S. K. Singh, S. Chejara, A. Gorovtsov et al. 2021. Sustainable approach and
safe use of biochar and its possible consequences. Sustainability. 13: 10362.
Fungal Strategies for the Remediation of Polycyclic Aromatic Hydrocarbons 105
John, W. C., I. O. Ogbonna, G. M. Gberikon and C. C. Iheukwumere. 2021. Evaluation of biosurfactant production
potential of Lysinibacillus fusiformis MK559526 isolated from automobile-mechanic-workshop soil. Braz. J.
Microbiol. 52: 663–674.
Jove, P., M. À. Olivella, S. Camarero, J. Caixach, C. Planas, L. Cano and F. X. De Las Heras. 2016. Fungal
biodegradation of anthracene-polluted cork: a comparative study. J. Environ. Sci. Health. Part A, 51(1): 70–77.
Kadri, T., T. Rouissi, S. K. Brar, M. Cledon, S. Sarma and M. Verma. 2017. Biodegradation of polycyclic aromatic
hydrocarbons (PAHs) by fungal enzymes: a review. J. Environ. Sci. 51: 52–74.
Kannangara, S., P. Ambadeniya, L. Undugoda and K. Abeywickrama. 2016. Polyaromatic hydrocarbon degradation
of moss endophytic fungi isolated from Macromitrium sp. in Sri Lanka. J. Agric. Sci. Technol. 6(03): 171–182.
Karigar, C. S. and S. S. Rao. 2011. Role of microbial enzymes in the bioremediation of pollutants: a review. Enzyme
Res.
Kariyawasam, T., G. S. Doran, J. A. Howitt and P. D. Prenzler. 2021. Polycyclic aromatic hydrocarbon contamination
in soils and sediments: sustainable approaches for extraction and remediation. Chemosphere. 132981.
Karlapudi, A. P., T. C. Venkateswarulu, J. Tammineedi, L. Kanumuri and B. K. Ravuru. 2018. Role of biosurfactants
in bioremediation of oil pollutiona review. Petroleum, 241–249.
Koukkou, A. I. and C. Drainas. 2008. Addressing PAH biodegradation in Greece: biochemical and molecular
approaches. IUBMB life. 60(5): 275–280.
Kour, D., T. Kaur, R. Devi, A. Yadav, M. Singh, D. Joshi and A. K. Saxena. 2021. Beneficial microbiomes for
bioremediation of diverse contaminated environments for environmental sustainability: present status and
future challenges. Environ. Sci. Pollut. Res. 28(20): 24917–24939.
Kuppusamy, S., P. Thavamani, K. Venkateswarlu, Y. B. Lee, R. Naidu and M. Megharaj. 2017. Remediation approaches
for polycyclic aromatic hydrocarbons (PAHs) contaminated soils: Technological constraints, emerging trends
and future directions. Chemosphere. 168: 944–968.
Langenbach, T. 2013. Persistence and bioaccumulation of persistent organic pollutants (POPs). pp. 305–329. In: Patil,
Y. B. and P. Rao [ed.]. Applied bioremediation—active and passive approaches. DOI: 10.5772/56418.
Lee, H., Y. Jang, Y. S. Choi, M. J. Kim, J. Lee, H. Lee et al. 2014. Biotechnological procedures to select white rot
fungi for the degradation of PAHs. J. Microbiol. Methods. 97: 56–62.
Lee, H., S. Y. Yun, S. Jang, G. H. Kim and J. J. Kim. 2015. Bioremediation of polycyclic aromatic hydrocarbons in
creosote-contaminated soil by Peniophora incarnata KUC8836. Bioremediat. J. 19(1): 1–8.
Leonardi, V., V. Sasek, M. Petruccioli, A. D’Annibale, P. Erbanová and T. Cajthaml. 2007. Bioavailability modification
and fungal biodegradation of PAHs in aged industrial soils. Int. Biodeterior. Biodegradation. 60(3): 165–170.
Li, X. and P. Li. 2008. Biodegradation of aged polycyclic aromatic hydrocarbons (PAHs) by microbial consortia in
soil and slurry phases. J. Hazard. Mater. 150(1): 21–26.
Li, X., Y. Wang, S. Wu, L. Qiu, L. Gu, J. Li et al. 2014. Peculiarities of metabolism of anthracene and pyrene by
laccase‐producing fungus Pycnoporus sanguineus H 1. Biotechnol. Appl. Biochem. 61(5): 549–554.
Li, X., Y. Pan, S. Hu, Y. Cheng, Y. Wang, K. Wu and S. Yang. 2018. Diversity of phenanthrene and benz [a] anthracene
metabolic pathways in white rot fungus Pycnoporus sanguineus 14. Int. Biodeterior. Biodegradation.
134: 25–30.
Lin, S., J. Wei, B. Yang, M. Zhang and R. Zhuo. 2022. Bioremediation of organic pollutants by white rot fungal
cytochrome P450: The role and mechanism of CYP450 in biodegradation. Chemosphere. 134776.
Lisowska, K. and J. Długoński. 2003. Concurrent corticosteroid and phenanthrene transformation by filamentous
fungus Cunninghamella elegans. J. Steroid Biochem. Mol. Biol. 85(1): 63–69.
Maia, M., A. Capao and L. Procópio, 2019. Biosurfactant produced by oil-degrading Pseudomonas putida AM-b1
strain with potential for microbial enhanced oil recovery. Bioremediat J. 23(4): 302–310.
Majcherczyk, A., C. Johannes and A. Hüttermann. 1998. Oxidation of polycyclic aromatic hydrocarbons (PAH) by
laccase of Trametes versicolor. Enzyme Microb. Technol. 22(5): 335–341.
Mandree, P., W. Masika, J. Naicker, G. Moonsamy, S. Ramchuran and R. Lalloo. 2021. Bioremediation of polycyclic
aromatic hydrocarbons from industry contaminated soil using indigenous Bacillus spp. Processes. 9(9): 1606.
Marco-Urrea, E., I. García-Romera and E. Aranda. 2015. Potential of non-ligninolytic fungi in bioremediation of
chlorinated and polycyclic aromatic hydrocarbons. N. Biotechnol. 32(6): 620–628.
Masih, J., R. Singhvi, K. Kumar, V. K. Jain and A. Taneja. 2012. Seasonal variation and sources of polycyclic
aromatic hydrocarbons (PAHs) in indoor and outdoor air in a semi-arid tract of northern India. Aerosol Air
Qual. Res. 12(4): 515–525.
McClements, D. J. and C. E. Gumus. 2016. Natural emulsifiers—Biosurfactants, phospholipids, biopolymers, and
colloidal particles: molecular and physicochemical basis of functional performance. Adv. Colloid Interface
Sci. 234: 3–26.
106 Bioremediation for Sustainable Environmental Cleanup
Premnath, N., K. Mohanrasu, R. G. R. Rao, G. H. Dinesh, G. S. Prakash, V. Ananthi and A. Arun. 2021. A crucial
review on polycyclic aromatic Hydrocarbons-Environmental occurrence and strategies for microbial
degradation. Chemosphere. 280: 130608.
Punnapayak, H., S. Prasongsuk, K. Messner, K. Danmek and P. Lotrakul. 2009. Polycyclic aromatic hydrocarbons
(PAHs) degradation by laccase from a tropical white rot fungus Ganoderma lucidum. Afr. J. Biotechnol. 8(21).
Rafin, C., O. Potin, E. Veignie, L. H. Sahraoui and M. Sancholle. 2000. Degradation of benzo [a] pyrene as sole
carbon source by a non-white rot fungus, Fusarium solani. Polycycl. Aromat. Comp. 21(1/4): 311–330.
Ravelet, C., S. Krivobok, L. Sage and R. Steiman. 2000. Biodegradation of pyrene by sediment
fungi. Chemosphere. 40(5): 557–563.
Ritz, K. and I. M. Young. 2004. Interactions between soil structure and fungi. Mycologist. 18(2): 52–59.
Romero, M. C., M. I. Urrutia, H. E. Reinoso and M. M. Kiernan. 2010. Benzo [a] pyrene degradation by soil
filamentous fungi. J. Yeast Fungal Res. 1(2): 025–029.
Rosales, E., M. Pazos and M. Ángeles Sanromán. 2013. Feasibility of solid‐state fermentation using spent fungi‐
substrate in the biodegradation of PAHs. CLEAN–Soil, Air, Water. 41(6): 610–615.
Ruiz-Duenas, F. J., M. Morales, M. Pérez-Boada, T. Choinowski, M. J. Martínez, K. Piontek and Á. T. Martínez.
2007. Manganese oxidation site in Pleurotus eryngii versatile peroxidase: a site-directed mutagenesis, kinetic,
and crystallographic study. Biochemistry. 46(1): 66–77.
Sack, U., M. Hofrichter and W. Fritsche. 1997. Degradation of polycyclic aromatic hydrocarbons by manganese
peroxidase of Nematoloma frowardii. FEMS Microbiol. Lett. 152(2): 227–234.
Saiu, G., S. Tronci, M. Grosso, E. Cadoni and N. Curreli. 2016. Biodegradation of polycyclic aromatic hydrocarbons
by pleurotus sajor-caju. Chem. Eng. Trans. 49: 487–492.
Saraswathy, A. and R. Hallberg. 2002. Degradation of pyrene by indigenous fungi from a former gasworks site. FEMS
Microbiol. Lett. 210(2): 227–232.
Saraswathy, A. and R. Hallberg. 2005. Mycelial pellet formation by Penicillium ochrochloron species due to exposure
to pyrene. Microbiol. Res. 160(4): 375–383.
Schmidt, S. N., J. H. Christensen and A. R. Johnsen. 2010. Fungal PAH-metabolites resist mineralization by soil
microorganisms. Environ. Sci. Technol. 44(5): 1677–1682.
Show, B. K., S. Banerjee, A. Banerjee, R. GhoshThakur, A. K. Hazra, N. C. Mandal and S. Chaudhury. 2022. Insect
gut bacteria: a promising tool for enhanced biogas production. Rev. Environ. Sci. Biotechnol., 1–25.
Sigoillot, J. C., J. G. Berrin, M. Bey, L. Lesage-Meessen, A. Levasseur, A. Lomascolo and E. Uzan-Boukhris. 2012.
Fungal strategies for lignin degradation. Adv. Bot. Res., 263–308.
Silva, I. S., M. Grossman and L. R. Durrant. 2009. Degradation of polycyclic aromatic hydrocarbons (2–7 rings) under
microaerobic and very-low-oxygen conditions by soil fungi. Int. Biodeterior. Biodegrad. 63(2): 224–229.
Sipilä, T. P., P. Väisänen, L. Paulin and K. Yrjälä. 2010. Sphingobium sp. HV3 degrades both herbicides and
polyaromatic hydrocarbons using ortho-and meta-pathways with differential expression shown by RT-PCR.
Biodegradation. 21(5): 771–784.
Srivastava, S. and M. Kumar. 2019. Biodegradation of polycyclic aromatic hydrocarbons (PAHs): a sustainable
approach. In Sustainable Green Technologies for Environmental Management. pp. 111–139. Springer,
Singapore.
Steffen, K. T., M. Hofrichter and A. Hatakka. 2002. Purification and characterization of manganese peroxidases from
the litter-decomposing basidiomycetes Agrocybe praecox and Stropharia coronilla. Enzyme Microb. Technol.
30(4): 550–555.
Sutherland, J. B. 1992. Detoxification of polycyclic aromatic hydrocarbons by fungi. J. Ind. Microbiol. Biotech. 9(1):
53–61.
Syed, K., H. Doddapaneni, V. Subramanian, Y. W. Lam and J. S. Yadav. 2010. Genome-to-function
characterization of novel fungal P450 monooxygenases oxidizing polycyclic aromatic hydrocarbons
(PAHs). Biochem. Biophys. Res. Commun. 399(4): 492–497.
Syed, K., A. Porollo, Y. W. Lam and J. S. Yadav. 2011. A fungal P450 (CYP5136A3) capable of oxidizing polycyclic
aromatic hydrocarbons and endocrine disrupting alkylphenols: role of Trp129 and Leu324. PloS one. 6(12):
28286.
Syed, K., A. Porollo, Y. W. Lam, P. E. Grimmett and J. S. Yadav. 2013. CYP63A2, a catalytically versatile fungal
P450 monooxygenase capable of oxidizing higher-molecular-weight polycyclic aromatic hydrocarbons,
alkylphenols, and alkanes. Appl. Environ. Microbiol. 79(8): 2692–2702.
Tavares, A. P. M., M. A. Z. Coelho, M. S. M. Agapito, J. A. P. Coutinho and A. Xavier. 2006. Optimization and
modeling of laccase production by Trametes versicolor in a bioreactor using statistical experimental design.
Appl. Biochem. Biotechnol. 134: 233–248.
108 Bioremediation for Sustainable Environmental Cleanup
Teerapatsakul, C., C. Pothiratana, L. Chitradon and S. Thachepan. 2016. Biodegradation of polycyclic aromatic
hydrocarbons by a thermotolerant white rot fungus Trametes polyzona RYNF13. J. Gen. Appl. Microbiol. 62(6):
303–312.
Ten Have, R. and P. J. M. Teunissen. 2001. Oxidative mechanisms 1952 involved in lignin degradation by white-rot
fungi. Chem. Rev. 1953 101: 3397–3414.
Thurston, C. F. 1994. The structure and function of fungal laccases. Microbiology. 140(1): 19–26.
Treu, R. and J. Falandysz. 2017. Mycoremediation of hydrocarbons with basidiomycetes—a review. J. Environ. Sci.
Health B. 52(3): 148–155.
Tripathi, V., S. A. Edrisi, B. Chen, V. K. Gupta, R. Vilu, N. Gathergood and P. C. Abhilash. 2017. Biotechnological
advances for restoring degraded land for sustainable development. Trends Biotechnol. 35(9): 847–859.
USEPA. 2002. Polycyclic Oganic Matter, US Environmental Protection Agency. (http://www.epa.gov/ttn/atw/hlthef/
polycycl.html). Accessed 06 May 2009.
Vasconcelos, M. R. S., G. A. L. Vieira, I. V. R. Otero, R. C. Bonugli-Santos, M. V. N. Rodrigues, V. L. G. Rehder
et al. 2019. Pyrene degradation by marine-derived ascomycete: process optimization, toxicity, and metabolic
analyses. Environ. Sci. Pollut. Res. 26: 12412–12424. https://doi.org/ 10.1007/s11356-019-04518-2.
Vipotnik, Z., M. Michelin and T. Tavares. 2022. Biodegradation of chrysene and benzo [a] pyrene and removal
of metals from naturally contaminated soil by isolated Trametes versicolor strain and laccase produced
thereof. Environ. Technol. Innov. 28: 102737.
Vipotnik, Z., M. Michelin and T. Tavares. 2021. Ligninolytic enzymes production during polycyclic aromatic
hydrocarbons degradation: effect of soil pH, soil amendments and fungal co-cultivation. Biodegradation.
32(2): 193–215.
Vyas, B. R. M., S. Bakowski, V. Sasek and M. Matucha. 1994. Degradation of anthracene by selected white-rot fungi.
FEMS Microbiol. Ecol. 14: 65–70.
Wang, R. Y., J. X. Liu, H. L. Huang, Z. Yu, X. M. Xu and G. M. Zeng. 2008. Effect of rhamnolipid on the enzyme
production of two species of lignin-degrading fungi. J. Hunan Univ. Nat. Sci. 35(10): 70–74.
Wang, Y., R. Vazquez-Duhalt and M. A. Pickard. 2003. Manganese–lignin peroxidase hybrid from Bjerkandera adusta
oxidizes polycyclic aromatic hydrocarbons more actively in the absence of manganese. Can. J. Microbiol. 49(11):
675–682.
Wong, D. W. 2009. Structure and action mechanism of ligninolytic enzymes. Appl. Biochem. Biotechnol. 157(2):
174–209.
Weis, L. M., A. M. Rummel, S. J. Masten, J. E. Trosko and B. L. Upham. 1998. Bay or baylike regions of polycyclic
aromatic hydrocarbons were potent inhibitors of Gap junctional intercellular communication. Environ. Health
Perspect. 106(1): 17–22.
Winquist, E., K. Björklöf, E. Schultz, M. Räsänen, K. Salonen, F. Anasonye et al. 2014. Bioremediation of PAH-
contaminated soil with fungi–from laboratory to field scale. Int. Biodeterior. Biodegradation. 86: 238–247.
Young, D., J. Rice, R. Martin, E. Lindquist, A. Lipzen, I. Grigoriev and D. Hibbett. 2015. Degradation of bunker
C fuel oil by white-rot fungi in sawdust cultures suggests potential applications in bioremediation. PloS
one. 10(6): 0130381.
Yuliani, H., M. S. Perdani, I. Savitri, M. Manurung, M. Sahlan, A. Wijanarko and H. Hermansyah. 2018. Antimicrobial
activity of biosurfactant derived from Bacillus subtilis C19. Energy Procedia. 153: 274–278.
Zhang, H., S. Zhang, F. He, X. Qin, X. Zhang and Y. Yang. 2016. Characterization of a manganese peroxidase from
white-rot fungus Trametes sp. 48424 with strong ability of degrading different types of dyes and polycyclic
aromatic hydrocarbons. J. Hazard. Mater. 320: 265–277.
Zhang, S., Y. Ning, X. Zhang, Y. Zhao, X. Yang, K. Wu et al. 2015. Contrasting characteristics of anthracene and
pyrene degradation by wood rot fungus Pycnoporus sanguineus H1. Int. Biodeterior. Biodegrad. 105: 228–232.
Zhang, X., X. Wang, C. Li, L. Zhang, G. Ning, W. Shi and Z. Yang. 2020. Ligninolytic enzyme involved in removal
of high molecular weight polycyclic aromatic hydrocarbons by Fusarium strain ZH-H2. Environ. Sci. Pollut.
Res. 27(34): 42969–42978.
Zhang, Y., X. Xiao, X. Zhu and B. Chen. 2022. Self-assembled fungus-biochar composite pellets (FBPs) for enhanced
co-sorption-biodegradation towards phenanthrene. Chemosphere. 286: 131887.
Zhuo, R. and F. Fan. 2021. A comprehensive insight into the application of white rot fungi and their lignocellulolytic
enzymes in the removal of organic pollutants. Sci. Total Environ. 778: 146132.
Zimmerman, A. R., D. H. Kang, M. Y. Ahn, S. Hyun and M. K. Banks. 2008. Influence of a soil enzyme on iron-
cyanide complex speciation and mineral adsorption. Chemosphere. 70(6): 1044–1051.
Chapter 7
Microbe-Assisted
Bioremediation of Pesticides
from Contaminated Habitats
Current Status and Prospects
Karen Reddy,1 Shisy Jose,1 Tufail Fayaz,2
Nirmal Renuka,2,* Sachitra Kumar Ratha,3
Sheena Kumari 1 and Faizal Bux 1
7.1 Introduction
The rapid growth of the global pesticide market, driven by widespread use of pesticides in
agriculture and non-agriculture sectors, has led to the introduction of numerous pesticide residues
into the environment (Mali et al. 2022). Pesticides are recalcitrant and non-biodegradable, thus when
applied to farmlands, gardens and other vegetation, they often remain toxic for years (Gonçalves
and Delabona 2022). They are also carcinogenic in nature and are banned in many countries due to
the risk created by their presence in the environment (Singh et al. 2020). In addition to polluting soil
and crops, they also pose a threat to ground water and other aquatic environments (Castelo-Grande
et al. 2010, Lehmann et al. 2018). Most pesticides reach destinations other than their intended target,
even though each is designed to eliminate a specific pest (Huang et al. 2018, Mali et al. 2022).
Several technologies have been developed and applied to contaminated sites to eliminate
the adverse environmental effects of pesticides. These include, physical treatments (adsorption
and percolator filters) and chemical treatments (advanced oxidation) (Satish et al. 2017). Even
though these methods seem promising, they are not cost effective and have several disadvantages.
As a result, a strategic plan for reducing agrochemical use and implementing sustainable farming
practices is essential (Gupta et al. 2016, Sun et al. 2020, Avila et al. 2021).
One such avenue that has been explored is bioremediation and degradation combined with
microbes. It is a cost effective and environmentally friendly technology of soil and water reclamation
1
Institute for Water and Wastewater Technology, Durban University of Technology, Durban, South Africa-4001.
2
Algal Biotechnology Laboratory, Department of Botany, Central University of Punjab, Bathinda, India-151401.
3
Phycology Laboratory, CSIR-National Botanical Research Institute, Lucknow, India-226001.
* Corresponding author: [email protected]
110 Bioremediation for Sustainable Environmental Cleanup
(Subashchandrabose et al. 2013). The two processes use living organisms to metabolize xenobiotics,
recalcitrant and toxic materials found in soil, water and sediments (Avila et al. 2021). There has been
extensive research on the biological removal of pesticides by microorganisms, including bacteria
(Arthrobacter, Bacillus, Corynebacterium, Flavobacterium, Pseudomonas and Rhodococcus), algae/
cyanobacteria (Chlorella, Scenedesmus, Synechococcus, Phormidium, Nostoc sp., Oscillatoria,
Anabaena sp., and Aulosira) and fungi (Penicillium, Aspergillus, Fusarium and Trichoderma)
(Nicolopoulou-Stamati et al. 2016). Several species display a strong pesticide degradation activity
and are highly adaptable (Subashchandrabose et al. 2013, Huang et al. 2018, Tarla et al. 2020). With
advanced metabolic mechanisms, microorganisms can use or transform pesticide molecules into
non-toxic or less hazardous forms of metabolites (Huang et al. 2018). This chapter aims to describe
the current status of pesticide pollution, and the types and mechanisms involved in microbe-assisted
bioremediation of pesticides.
7.2.3 Carbamates
Carbamate compounds are commonly used as insecticides (esters of carbamic acid). While, the
derivatives of carbamic acid, thiocarbamic acid and dithiocarbamic acid are used as herbicides
(Gupta 2014). Regularly known carbamate pesticides are aldicarb, bendiocarb, carbofuran,
carbosulfan, carbaryl, methomyl, oxamyl, propoxur and ziram (Gupta 2014, Huang et al. 2018).
In soil and water, carbamates tend to hydrolyze easily, thus resulting in low levels of persistence in
the environment. The direct application of carbamates in soil can cause a significant reduction in
the microflora, worms and some of the compounds are toxic to mammals and birds (Basheer et al.
2009). Additionally, the excessive use of carbamates can disrupt enzymes that regulate acetylcholine,
which is a neurotransmitter in the nervous system. They cause disorders in reproductive and cellular
metabolic functions in animals and humans (Huang et al. 2018, Gonçalves and Delabona 2022).
112 Bioremediation for Sustainable Environmental Cleanup
Figure 7.1. The top 30 countries susceptible to high pesticide pollution risk (Tang et al. 2021).
A bioremediation method can be classified into three subcategories: in-situ, ex-situ solid or ex-situ
slurry. In-situ techniques treat soils and associated groundwater on-site without excavation, while
ex-situ techniques require excavation prior to treatment (Shukla et al. 2010). Three fundamental
principles guide the choice of an appropriate bioremediation technology, namely the ability of the
pollutant to be transformed biologically, the accessibility of the contaminant to microorganisms and
the possibility of optimizing biological activity (Dua et al. 2002). Biological activity is achieved as
microorganisms use the contaminants as nutrients or energy sources. By introducing microorganisms
with desired catalytic capabilities or by supplementing nutrients (nitrogen and phosphorus), electron
acceptors (oxygen) and substrates (methane, phenol, and toluene), the activity of the microbe can
be further stimulated (Shukla et al. 2010). However, the appropriate degrading enzymes and several
environmental factors also contribute to the remediation of pesticides. The possible roles and
applications of microbes for the remediation of pesticides are further summarized.
Table 7.2. The remediation of pesticides by common bacterial strains from the environment.
bioremediation of diverse chemical classes (Singh et al. 2020). Pesticides are remediated by bacteria
based on species specificity and several abiotic factors, such as temperature, pH, nutrient content,
moisture and humidity (Huang et al. 2018). Among pesticides, bacteria are the main degraders
of organochlorines. These are synthetic organic compounds containing at least one covalently
bonded chlorine atom and are insecticides primarily composed of carbon, hydrogen and chlorine.
Some of the most known organochlorine pesticides include, dieldrin, aldrin, lindane, endosulfan,
dichlorodiphenyltrichloroethane (DDT) and hexachlorocyclohexane (HCH). Pesticides containing
organophosphates (imidacloprid, diazinon and chlorpyrifos) are also among the chemical groups
investigated for degradation by bacteria (Jayaraj et al. 2016).
A few earlier reports (Table 7.2) had suggested the potential of Pseudomonas sp. for the
remediation of insecticides and herbicides. Endosulfan (insecticide) was bioremediated by
Pseudomonas fluorescens during a 5 d laboratory study. The results of the study showed that up to
80% of endosulfan could be remediated (Zaffar et al. 2018). In another study, approximately 99.9%
of atrazine (herbicide) was remediated after 2 d incubation with Pseudomonas sp. There has also been
research into the use of bacterial mixtures for pesticide remediation (Cai et al. 2003). When a mixture
of Pseudomonas stutzeri, Pseudomonas aeruginosa and Bacillus firmus was used in a week-long
laboratory study, approximately 68% of the DDT (insecticide) was remediated. Imidacloprid, another
insecticide, was remediated by Klebsiella pneumonia (78%) and Bacillus subtilis (25.36–45.48%)
during 7-d and 25-d experiments, respectively (Phugare et al. 2013, Sabourmoghaddam et al. 2015).
Additionally, Burkholderia and Acinetobacter were found to be capable of remediating pesticides
during tests conducted over a week. Almost all the potential bacterial strains were isolated from soil
samples, indicating the cost-effectiveness of using indigenous bacteria for pesticide remediation.
Bacteria can interact, both chemically and physically, with substances, leading to structural
changes or complete degradation of the target molecule (Ortiz-Hernández et al. 2013). Bacteria
can transform or degrade pesticides into less toxic or non-toxic forms (McGuinness and Dowling
2009). This is commonly known as a detoxification mechanism (Figure 7.2), where bacteria produce
intracellular and extracellular enzymes (Singh et al. 2020). The activity of enzymes depends on the
metabolic potential of the bacteria to detoxify or transform the pollutants, which depends on both
accessibility and bioavailability (Ramakrishnan et al. 2011). There could be three phases to the
metabolism of pesticides. Phase one involves the transformation of the initial properties of the parent
compound through oxidation, reduction or hydrolysis. This transformation produces a more water-
soluble and usually less toxic product than the parent. In the second phase, a pesticide or pesticide
Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats 115
metabolite is conjugated to a sugar or amino acid, that increases its water solubility and reduces its
toxicity. During the third phase, phase two metabolites are converted into secondary conjugates,
which are also non-toxic (Ramakrishnan et al. 2011). The overall metabolism of pesticides by
enzymes proves to be favorable, as enzymes are more resistant to environmental conditions, which
makes it easier for them to remove pesticides more efficiently from affected areas (Huang et al.
2008). Therefore, it is imperative to identify species of bacteria that can thrive in extreme conditions
and completely degrade pesticides. A combination of mechanistic approaches and microbial genetic
advancement can enhance sustainable pesticide bioremediation by identifying the most effective
bacterial strains or bacterial consortia.
d – days
In the biosorption process, the liquid and solid phases contain the dissolved or suspended pesticides
to be absorbed. It could be defined as the attachment of potentially toxic pesticides to the surface
of the photosynthetic strain. Pesticides are biosorbed in a passive and metabolically independent
process that occurs faster than bioaccumulation (Verasoundarapandian et al. 2022). Strains
such as Chlorella and Fischerella have been reported to remediate > 80% of pesticides through
biosorption (Hussein et al. 2017, Tiwari et al. 2017). While bioaccumulation studies on fluroxypyr,
prometryne and atazarine, have shown that members of Chlorophyta are able to accumulate
pesticides during 4–5-d intervals but at a lower rate (< 60%) (Jin et al. 2012, Kabra et al. 2014,
Zhang et al. 2021). The bioaccumulation ability of organisms such as algae is determined by their
lipid content, which is influenced by their growth conditions and cell distribution (Sakurai et al.
2016). In addition to bioaccumulating pesticides, these photosynthetic organisms are also able to
transform and degrade pesticides into non-toxic or less toxic compounds. Pesticide degradation is
influenced by microorganism types, optimal environmental conditions and the metabolic activity of
various enzymes (hydrolase, phosphatase, phosphodiesterase, oxygenase, esterase, transferase and
oxidoreductases) (Verasoundarapandian et al. 2022). The probable biosorption, bioaccumulation
and biodegradation processes for the removal of pesticides in photosynthetic microorganisms are
shown in Figure 7.3.
Phycoremediation technology offers additional benefits such as minimized greenhouse
gas emissions from the environment and biomass reuse (Renuka et al. 2018, Reddy et al. 2021,
Renuka et al. 2021). Therefore, phycoremediation of pesticides can be a promising integrated and
sustainable approach for eco-friendly and efficient removal of pesticides from contaminated areas
with additional environmental benefits (Singh et al. 2020).
Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats 117
Irprex lacteus demonstrated degradation rates greater than 70%, when grown in the presence of
chlorpyrifos at a temperature of 30°C and pH 7 (Wang et al. 2020). In another study, Fang et al.
(2008), reported the removal of chlorpyrifos by the fungal strain verticillium sp. with high removal
efficiency. Kulshrestha and Kumari (2011), explored the potential removal of chlorpyrifos in both
mixed and pure fungi. They reported that mixed cultures could remove up to 300 mg L–1 chlorpyrifos.
Another widely used insecticide of interest, allethrin, has a direct impact on human health. Bhatt
et al. (2020) isolated Fusarium proliferatum CF2 from an allethrin contaminated agriculture site.
The isolate could completely degrade the allethrin in 6 d. Endosulfan is degraded by a variety of
fungal strains, including Paecilomyces variotii, Paecilomyces lilacinus (Hernández-Ramos et al.
2019), Aspergillus sydoni (Goswami et al. 2009), Pleurotus eryngi (Wang et al. 2018), Aspergillus
niger, Penicillium chrysogenum and Aspergillus flavus (Ahmad 2020). Endosulfan degradation
was also reported using Trametes hirsute (Kamei et al. 2011). Two pathways were reported, i.e.,
hydrolysis and oxidation of endosulfan that convert it to endosulfan diol and endosulfan sulfate,
respectively. In an aqueous medium and soil, the bacterial and fungal consortia could remove
endosulfan effectively (Abraham and Silambarasan 2014).
Several studies have reported the efficacy of fungi to degrade different pesticides (Table 7.4).
Phanerochaete chrysosporium grown in 10 mg L–1 of thiamethoxam resulted in degradation rates
of 49 and 98% during 15 and 25 d, respectively (Chen et al. 2021). An estimated 71% of triclosan
was degraded by Aspergillus versicolor (Ertit Taştan and Dönmez 2015). Trichoderma harzianum
CBMAI 167 isolated from a marine environment was grown in 50 mg L–1 of pentachlorophenol
(PCP), which resulted in complete degradation to pentachloroanisole (PCA) after incubation for
7 d (Vacondio et al. 2015). It was also observed that PCA was degraded by Trichoderma harzianum
CBMAI 167. When Bjerkandera adusta and Anthracophyllum discolor species were grown in PCP
contaminated soil (100, 250 and 350 mg kg–1), both fungi degraded PCP, however, the highest
Table 7.4. Pesticide removal efficiency of various fungal strains from the environment.
d – days
Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats 119
Acknowledgment
The authors hereby acknowledge the Durban University of Technology, South Africa (UID: 84166),
Central University of Punjab, Bathinda, India (CUPB/Acad./2022/1194), and CSIR-National
Botanical Research Institute, Lucknow, India for providing facilities.
120 Bioremediation for Sustainable Environmental Cleanup
References
Abdel-Fattah Mostafa, A., M. T. Yassin, T. M. Dawoud, F. O. Al-Otibi and S. R. M. Sayed. 2022. Mycodegradation of
diazinon pesticide utilizing fungal strains isolated from polluted soil. Environ. Res. 212: 113421.
Abraham, J. and S. Silambarasan. 2014. Biomineralization and formulation of endosulfan degrading bacterial and
fungal consortiums. Pestic. Biochem. Physiol. 116: 24–31.
Ahmad, K. S. 2020. Remedial potential of bacterial and fungal strains (Bacillus subtilis, Aspergillus niger, Aspergillus
flavus and Penicillium chrysogenum) against organochlorine insecticide Endosulfan. Folia Microbiol. (Praha).
65(5): 801–810.
Ajaz, M., S. A. Rasool, S. K. Sherwani and T. A. Ali. 2012. High profile chlorpyrifos degrading pseudomonas putida
mas-1 from indigenous soil: gas chromatographic analysis and molecular characterization. Int. J. Basic Appl.
Med. Sci. (IJBMSP). 2(2).
Alamgir Zaman Chowdhury, M., A. N. M. Fakhruddin, M. Nazrul Islam, M. Moniruzzaman, S. H. Gan and
M. Khorshed Alam. 2013. Detection of the residues of nineteen pesticides in fresh vegetable samples using gas
chromatography–mass spectrometry. Food Control. 34(2): 457–465.
Alvarenga, N., W. G. Birolli, M. H. Seleghim and A. L. Porto 2014. Biodegradation of methyl parathion by whole
cells of marine-derived fungi Aspergillus sydowii and Penicillium decaturense. Chemosphere. 117: 47–52.
Alvarenga, N., W. G. Birolli, M. Nitschke, M. O. de O Rezende, M. Seleghim and A. Porto. 2015. Biodegradation
of chlorpyrifos by whole cells of marine-derived fungi Aspergillus sydowii and Trichoderma sp. J. Microb.
Biochem. Technol. 7: 133–139.
Amani, F., A. A. Safari Sinegani, F. Ebrahimi and S. Nazarian. 2018. Biodegradation of chlorpyrifos and diazinon
organophosphates by two bacteria isolated from contaminated agricultural soils. Biol. J. Microorganism.
7(28): 27–39.
Avila, R., A. Peris, E. Eljarrat, T. Vicent and P. Blánquez. 2021. Biodegradation of hydrophobic pesticides by
microalgae: transformation products and impact on algae biochemical methane potential. Sci. Total Environ.
754: 142114.
Azubuike, C. C., C. B Chikere and G. C. Okpokwasili. 2016. Bioremediation techniques–classification based on site
of application: principles, advantages, limitations and prospects. World J. Microbiol. Biotechnol. 32(11): 180.
Bal, R.., M. Naziroğlu, G. Türk, Ö. Yilmaz, T. Kuloğlu, E. Etem and G. Baydas. 2012. Insecticide imidacloprid
induces morphological and DNA damage through oxidative toxicity on the reproductive organs of developing
male rats. Cell Biochem. Funct. 30(6): 492–499.
Basheer, C., A. A. Alnedhary, B. S. M. Rao and H. K. Lee. 2009. Determination of carbamate pesticides using micro-
solid-phase extraction combined with high-performance liquid chromatography. J. Chromatogr. A. 1216(2):
211–216.
Bernardino, J., A. M. Bernardette, M. Ramrez-Sandoval and D. Domnguez-Oje. 2012. Biodegradation and
bioremediation of organic pesticides. pp. 253–272. In: R. P. Soundararajan [Ed.]. Pesticides: Recent Trends in
Pesticide Residue Assay. London: IntechOpen.
Bhalerao, T. S. and P. R. Puranik. 2007. Biodegradation of organochlorine pesticide, endosulfan, by a fungal soil
isolate, Aspergillus niger. Int. Biodeterior. Biodegradation. 59(4): 315–321.
Bhandari, G. 2017. Mycoremediation: an eco-friendly approach for degradation of pesticides. pp. 119–131.
In: R. Prasad [Ed.]. Mycoremediation and Environmental Sustainability. (1). Cham: Springer International
Publishing.
Bhatt, P., W. Zhang, Z. Lin, S. Pang, Y. Huang and S. Chen. 2020. Biodegradation of allethrin by a novel fungus
Fusarium proliferatum strain CF2, Isolated from Contaminated Soils. Microorganisms. 8(4): 593.
Bodin, H. 2014. Phycoremediation of pesticides using microalgae. MSc thesis, Swedish University of Agricultural
Sciences, Sweden.
Bose, S., P. S. Kumar and D.-V. N. Vo. 2021. A review on the microbial degradation of chlorpyrifos and its metabolite
TCP. Chemosphere. 283: 131447.
Cai, B., Y. Han, B. Liu, Y. Ren and S. Jiang. 2003. Isolation and characterization of an atrazine-degrading bacterium
from industrial wastewater in China. Lett. Appl. Microbiol. 36(5): 272–276.
Castelo-Grande, T., P. A. Augusto, P. Monteiro, A. M. Estevez and D. Barbosa. 2010. Remediation of soils
contaminated with pesticides: a review. Int. J. Environ. Anal. Chem. 90(3-6): 438–467.
Chen, A., G. Zeng, G. Chen, C. Zhang, M. Yan, C. Shang, X. Hu, L. Lu, M. Chen, Z. Guo and Y. Zuo. 2014. Hydrogen
sulfide alleviates 2,4-dichlorophenol toxicity and promotes its degradation in Phanerochaete chrysosporium.
Chemosphere. 109: 208–212.
Chen, A., W. Li, X. Zhang, C. Shang, S. Luo, R. Cao and D. Jin. 2021. Biodegradation and detoxification of
neonicotinoid insecticide thiamethoxam by white-rot fungus Phanerochaete chrysosporium. J. Hazard Mater.
417: 126017.
Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats 121
Chen, S., C. Liu, C. Peng, H. Liu, M. Hu and G. Zhong. 2012. Biodegradation of chlorpyrifos and its hydrolysis
product 3,5,6-trichloro-2-pyridinol by a new fungal strain Cladosporium cladosporioides Hu-01. PLoS One,
7(10): e47205.
DeClementi, C. and B. R. Sobczak. 2012. Common rodenticide toxicoses in small animals. Vet Clin. North Am. Small
Anim. Pract. 42(2): 349–360.
Degrendele, C., J. Klánová, R. Prokeš, P. Příbylová, P. Šenk, M. Šudoma, M. Röösli, M. A. Dalvie and S. Fuhrimann.
2022. Current use pesticides in soil and air from two agricultural sites in South Africa: implications for
environmental fate and human exposure. Sci. Total Environ. 807(Pt 1): 150455.
Deng, Y.-J. and Wang, S. Y. 2016. Synergistic growth in bacteria depends on substrate complexity. J. Microbiol. 54(1):
23–30.
Deshmukh, R., A. A. Khardenavis and H. J. Purohit. 2016. Diverse metabolic capacities of fungi for bioremediation.
Indian J. Microbiol. 56(3): 247–264.
Dosnon-Olette, R., P. Trotel-Aziz, M. Couderchet and P. Eullaffroy. 2010. Fungicides and herbicide removal in
Scenedesmus cell suspensions. Chemosphere. 79(2): 117–123.
Dua, M., A. Singh, N. Sethunathan and A. Johri. 2002. Biotechnology and bioremediation: successes and limitations.
Appl. Microbiol. Biotechnol. 59(2): 143–152.
Ertit Taştan, B. and G. Dönmez. 2015. Biodegradation of pesticide triclosan by A. versicolor in simulated wastewater
and semi-synthetic media. Pestic Biochem Physiol. 118: 33–37.
Fang, H., Y. Q. Xiang, Y. J. Hao, X. Q. Chu, X. D. Pan, J. Q. Yu and Y. L. Yu. 2008. Fungal degradation of chlorpyrifos
by Verticillium sp. DSP in pure cultures and its use in bioremediation of contaminated soil and pakchoi.
Int. Biodeterior. Biodegradation. 61(4): 294–303.
Foong, S. Y., N. L. Ma, S. S. Lam, W. Peng, F. Low, B. H. Lee, A. K. Alstrup and C. Sonne, 2020. A recent global
review of hazardous chlorpyrifos pesticide in fruit and vegetables: prevalence, remediation and actions needed.
J. Hazard. Mater. 400: 123006.
Gonçalves, C. R. and P. D. S. Delabona. 2022. Strategies for bioremediation of pesticides: challenges and perspectives
of the Brazilian scenario for global application – A review. Environ. Adv. 8: 100220.
Goswami, S., K. Vig and D. K. Singh. 2009. Biodegradation of α and β endosulfan by Aspergillus sydoni. Chemosphere.
75(7): 883–888.
Goulson, D. 2013. An overview of the environmental risks posed by neonicotinoid insecticides. J. Appl. Ecol. 50(4):
977–987.
Gupta, M., S. Mathur, T. K. Sharma, M. Rana, A. Gairola, N. K. Navani and R. Pathania. 2016. A study on metabolic
prowess of Pseudomonas sp. RPT 52 to degrade imidacloprid, endosulfan and coragen. J. Hazard Mater. 301:
250–258.
Gupta, R. C. 2014. Carbamate pesticides. pp. 661–664. In: P. Wexler [Ed.]. Encyclopedia of Toxicology (Third
Edition). Oxford. Academic Press.
Hajihassani, A., K. S. Lawrence and G. B. Jagdale. 2018. Plant parasitic nematodes in Georgia and Alabama Plant
parasitic nematodes in sustainable agriculture of North America. pp. 357–391. Springer.
Hamad, M. T. M. H. 2020. Biodegradation of diazinon by fungal strain Apergillus niger MK640786 using response
surface methodology. Environ. Technol. Innov. 18: 100691.
Harms, H., D. Schlosser and L. Y. Wick. 2011. Untapped potential: exploiting fungi in bioremediation of hazardous
chemicals. Nat. Rev. Microbiol. 9(3): 177–192.
Hernández-Ramos, A. C., S. Hernández and I. Ortíz. 2019. Study on endosulfan-degrading capability of Paecilomyces
variotii, Paecilomyces lilacinus and Sphingobacterium sp. in liquid cultures. Bioremed. J. 23(4): 251–258.
Hoshi, N., T. Hirano, T. Omotehara, J. Tokumoto, Y. Umemura, Y. Mantani, T. Tanida, K. Warita, Y. Tabuchi,
T. Yokoyama and H. Kitagawa. 2014. Insight into the mechanism of reproductive dysfunction caused by
neonicotinoid pesticides. Biol. Pharm. Bull. 37(9): 1439–1443.
Huang, D.-L., G.-M. Zeng, C.-L. Feng, S. Hu, X.-Y. Jiang, L. Tang, F. F. Su, Y. Zhang, W. Zeng and H. L. Liu. 2008.
Degradation of lead-contaminated lignocellulosic waste by Phanerochaete chrysosporium and the reduction
of lead toxicity. Environ. Sci. Technol. 42(13): 4946–4951.
Huang, Y., Xiao, L., Li, F., Xiao, M., Lin, D., Long, X. and Wu, Z. 2018. Microbial degradation of pesticide residues
and an emphasis on the degradation of cypermethrin and 3-phenoxy benzoic acid: a review. Molecules. 23(9):
2313.
Hussein, M., A. Abdullah, N. Badr El-Din and E. Mishaqa. 2017. Biosorption potential of the microchlorophyte
Chlorella vulgaris for some pesticides. J. Fertil. Pestic. 8(01): 1000177.
Ibrahim, W. M., M. A. Karam, R. M. El-Shahat and A. A. Adway. 2014. Biodegradation and utilization of
organophosphorus pesticide malathion by cyanobacteria. Biomed Res. Int. 392682.
Jayaraj, R., P. Megha and P. Sreedev. 2016. Organochlorine pesticides, their toxic effects on living organisms and their
fate in the environment. Interdiscip. Toxicol. 9(3-4): 90–100.
122 Bioremediation for Sustainable Environmental Cleanup
Jin, Z. P., K. Luo, S. Zhang, Q. Zheng and Yang, H. 2012. Bioaccumulation and catabolism of prometryne in green
algae. Chemosphere. 87(3): 278–284.
Kabra, A. N., M.-K. Ji, J. Choi, J. R. Kim, S. P. Govindwar and B.-H. Jeon. 2014. Toxicity of atrazine and its
bioaccumulation and biodegradation in a green microalga, Chlamydomonas mexicana. Environ. Sci. Pollut.
Res. 21(21): 12270–12278.
Kafilzadeh, F. 2015. Assessment of organochlorine pesticide residues in water, sediments and fish from Lake Tashk,
Iran. Achiev. Life Sci. 9(2): 107–111.
Kalyabina, V. P., E. N. Esimbekova, K. V. Kopylova and V. A. Kratasyuk. 2021. Pesticides: formulants, distribution
pathways and effects on human health—a review. Toxicol. Rep. 8: 1179–1192.
Kamei, I., K. Takagi and R. Kondo. 2011. Degradation of endosulfan and endosulfan sulfate by white-rot fungus
Trametes hirsuta. J. Wood Sci. 57(4): 317–322.
Kaur, H., S. Kapoor and G. Kaur. 2016. Application of ligninolytic potentials of a white-rot fungus Ganoderma
lucidum for degradation of lindane. Environ. Monit. Assess. 188(10): 588.
Kolaczinski, J. H. and C. F. Curtis 2004. Chronic illness as a result of low-level exposure to synthetic pyrethroid
insecticides: a review of the debate. Food Chem. Toxicol. 42(5): 697–706.
Kulshrestha, G. and A. Kumari. 2011. Fungal degradation of chlorpyrifos by Acremonium sp. strain (GFRC-1)
isolated from a laboratory-enriched red agricultural soil. Biol. Fertil. Soils. 47(2): 219–225.
Kumar, A., A. Sharma, P. Chaudhary and S. Gangola. 2021. Chlorpyrifos degradation using binary fungal strains
isolated from industrial waste soil. Biologia. 76(10): 3071–3080.
Landrigan, P. J., R. Fuller, N. J. Acosta, O. Adeyi, R. Arnold, A. B. Baldé, R. Bertollini, S. Bose-O’Reilly,
J. I. Boufford, P. N. Breysse and T. Chiles. 2018. The Lancet Commission on pollution and health. The Lancet.
391(10119): 462–512.
Lehmann, E., M. Fargues, J.-J. Nfon Dibié, Y. Konaté and L. F. de Alencastro. 2018. Assessment of water resource
contamination by pesticides in vegetable-producing areas in Burkina Faso. Environ. Sci. Pollut. Res. 25(4):
3681–3694.
Mali, H., C. Shah, B. H. Raghunandan, A. S. Prajapati, D. H. Patel, U. Trivedi and R. B. Subramanian. 2022.
Organophosphate pesticides an emerging environmental contaminant: pollution, toxicity, bioremediation
progress, and remaining challenges. J. Environ. Sci. 127: 234–250.
Mandal, K., B. Singh, M. Jariyal and V. K. Gupta. 2013. Microbial degradation of fipronil by Bacillus thuringiensis.
Ecotoxicol. Environ. Saf. 93: 87–92.
Mata, T. M., A. A. Martins and N. S. Caetano. 2010. Microalgae for biodiesel production and other applications: a
review. Renew. Sustain. Energy Rev. 14(1): 217–232.
Matsumoto, E., Y. Kawanaka, S. J. Yun and H. Oyaizu. 2008. Isolation of dieldrin- and endrin-degrading bacteria
using 1,2-epoxycyclohexane as a structural analog of both compounds. Appl. Microbiol. Biotechnol. 80(6):
1095–1103.
McGuinness, M. and D. Dowling. 2009. Plant-associated bacterial degradation of toxic organic compounds in soil.
Int. J. Environ. Res. Public Health. 6(8): 2226–2247.
Mir-Tutusaus, J. A., M. Masís-Mora, C. Corcellas, E. Eljarrat, D. Barceló, M. Sarrà, G. Caminal, T. Vicent and
C. E. Rodríguez-Rodríguez. 2014. Degradation of selected agrochemicals by the white rot fungus Trametes
versicolor. Sci. Total Environ. 500: 235–242.
Nicolopoulou-Stamati, P., S. Maipas, C. Kotampasi, P. Stamatis and L. Hens. 2016. Chemical pesticides and human
health: the urgent need for a new concept in agriculture. Public Health Front. 4: 148.
Nile, A. S., Kwon, Y. D. and Nile, S. H. 2019. Horticultural oils: possible alternatives to chemical pesticides and
insecticides. Environ. Sci. Pollut. Res. 26(21): 21127–21139.
Ortiz-Hernández, M. L., E. Sánchez-Salinas, E. Dantán-González and M. L. Castrejón-Godínez. 2013. Pesticide
biodegradation: mechanisms, genetics and strategies to enhance the process. Biodegradation-life of Science.
10: 251–287.
Peter, L., A. Gajendiran, D. Mani, S. Nagaraj and J. Abraham. 2015. Mineralization of malathion by Fusarium
oxysporum strain JASA1 isolated from sugarcane fields. Environ. Prog. Sustain. Energy. 34(1): 112–116.
Phugare, S. S., D. C. Kalyani, Y. B. Gaikwad and J. P. Jadhav. 2013. Microbial degradation of imidacloprid and
toxicological analysis of its biodegradation metabolites in silkworm (Bombyx mori). Chem. Eng. J. 230: 27–35.
Powthong, P., B. Jantrapanukorn and P. Suntornthiticharoen. 2016. Isolation, identification and analysis of DDT-
degrading bacteria for agriculture area improvements. J. Food Agric. Environ. 14(1): 131–136.
Rajmohan, K. S., R. Chandrasekaran and S. Varjani 2020. A Review on occurrence of pesticides in environment and
current technologies for their remediation and management. Indian J. Microbiol. 60(2): 125–138.
Ramakrishnan, B., M. Megharaj, K. Venkateswarlu, N. Sethunathan and R. Naidu 2011. Mixtures of environmental
pollutants: effects on microorganisms and their activities in soils. Rev. Environ. Contam. Toxicol. 211: 63–120.
Microbe-Assisted Bioremediation of Pesticides from Contaminated Habitats 123
Ray, D. E. and J. R. Fry. 2006. A reassessment of the neurotoxicity of pyrethroid insecticides. Pharmacol. Ther.
111(1): 174–193.
Reddy, K., N. Renuka, S. Kumari and F. Bux. 2021. Algae-mediated processes for the treatment of antiretroviral drugs
in wastewater: Prospects and challenges. Chemosphere. 280: 130674.
Renuka, N., A. Gulde, R. Prasanna, P. Singh and F. Bux. 2018. Microalgae as multi-functional options in modern
agriculture: current trends, prospects and challenges. Biotechnol. Adv. 36(4): 1255–1273.
Renuka, N., S. K. Ratha, F. Kader, I. Rawat and F. Bux. 2021. Insights into the potential impact of algae-mediated
wastewater beneficiation for the circular bioeconomy: a global perspective. J. Environ. Manage. 297: 113257.
Rodríguez-Rodríguez, C. E., K. Madrigal-León, M. Masís-Mora, M. Pérez-Villanueva and J. S. Chin-Pampillo. 2017.
Removal of carbamates and detoxification potential in a biomixture: fungal bioaugmentation versus traditional
use. Ecotoxicol. Environ. Saf. 135: 252–258.
Rubilar, O., G. Feijoo, C. Diez, T. A. Lu-Chau, M. T. Moreira and J. M. Lema. 2007. Biodegradation of
Pentachlorophenol in Soil Slurry Cultures by Bjerkandera adusta and Anthracophyllum discolor. Ind. Eng.
Chem. Res. 46(21): 6744–6751.
Sabourmoghaddam, N., M. P. Zakaria and D. Omar. 2015. Evidence for the microbial degradation of imidacloprid in
soils of Cameron Highlands. J. Saudi Soc. Agric. Sci. 14(2): 182–188.
Sadiq, S., M. Haq, I. Ahmad, K. Ahad, A. Rashid and N. Rafiq. 2015. Bioremediation potential of white rot fungi,
Pleurotus spp. against organochlorines. J. Bioremed. Biodeg. 6(308): 2.
Saez, J. M., A. L. Bigliardo, E. E. Raimondo, G. E. Briceño, M. A. Polti and C. S. Benimeli. 2018. Lindane dissipation
in a biomixture: effect of soil properties and bioaugmentation. Ecotoxicol. Environ. Saf. 156: 97–105.
Sagar, V. and D. Singh. 2011. Biodegradation of lindane pesticide by non white-rots soil fungus Fusarium sp. World
J. Microbiol. Biotechnol. 27(8): 1747–1754.
Sakurai, T., M. Aoki, X. Ju, T. Ueda, Y. Nakamura, S. Fujiwara, T. Umemura, M. Tsuzuki and A. Minoda. 2016.
Profiling of lipid and glycogen accumulations under different growth conditions in the sulfothermophilic red
alga Galdieria sulphuraria. Bioresour. Technol. 200: 861–866.
Sanchez, S. and A. L. Demain. 2017. Bioactive products from fungi. pp. 59–87. In: M. Puri [Ed.]. Food Bioactives.
Springer, Cham.
Sarker, A., R. Nandi, J.-E. Kim and T. Islam. 2021. Remediation of chemical pesticides from contaminated sites
through potential microorganisms and their functional enzymes: prospects and challenges. Environ. Technol.
Innov. 23: 101777.
Satish, G. P., D. M. Ashokrao and S. K. Arun. 2017. Microbial degradation of pesticide: a review. Afr. J. Microbiol.
Res. 11(24): 992–1012.
Shahi, A., S. Aydin, B. Ince and O. Ince. 2016. The effects of white-rot fungi Trametes versicolor and Bjerkandera
adusta on microbial community structure and functional genes during the bioaugmentation process following
biostimulation practice of petroleum contaminated soil. Int. Biodeterior. Biodegrad. 114: 67–74.
Shukla, K. P., N. K. Singh and S. Sharma 2010. Bioremediation: developments, current practices and perspectives.
Genet. Eng. Biotechnol. J. 3: 1–20.
Singh, B. K. and A. Walker. 2006. Microbial degradation of organophosphorus compounds. FEMS Microbiol. Rev.
30(3): 428–471.
Singh, D. V., R. Ali, M. Kulsum and R. A. Bhat. 2020. Ecofriendly approaches for remediation of pesticides in
contaminated environs. pp. 173–194. In: R. Bhat, K. Hakeem and N. Saud Al-Saud [Eds.]. Bioremediation and
Biotechnology, Vol 3. Springer, Cham.
Subashchandrabose, S. R., B. Ramakrishnan, M. Megharaj, K. Venkateswarlu and R. Naidu 2013. Mixotrophic
cyanobacteria and microalgae as distinctive biological agents for organic pollutant degradation. Environ. Int.
51: 59–72.
Sun, Y., M. Kumar, L. Wang, J. Gupta and D. C. W. Tsang. 2020. 13—Biotechnology for soil decontamination: opportunity,
challenges, and prospects for pesticide biodegradation. pp. 261–283. In: F. Pacheco-Torgal, V. Ivanov and
D. C. W. Tsang (Eds.). Bio-Based Materials and Biotechnologies for Eco-Efficient Construction: Woodhead
Publishing.
Tang, F. H. M., M. Lenzen, A. McBratney and F. Maggi. 2021. Risk of pesticide pollution at the global scale. Nat.
Geosci. 14(4): 206–210.
Tarla, D. N., L. E. Erickson, G. M. Hettiarachchi, S. I. Amadi, M. Galkaduwa, L. C. Davis, A. Nurzhanova and
V. Pidlisnyuk. 2020. Phytoremediation and bioremediation of pesticide-contaminated soil. Appl. Sci. 10(4):
1217.
Tiwari, B., S. Chakraborty, A. K. Srivastava and A. K. Mishra. 2017. Biodegradation and rapid removal of methyl
parathion by the paddy field cyanobacterium Fischerella sp. Algal Res. 25: 285–296.
124 Bioremediation for Sustainable Environmental Cleanup
8.1 Introduction
The world’s population is rapidly increasing. The current global population is estimated to be
7.97 billion people as of 2020. This has obvious consequences for the world in which we live.
To meet the needs of a growing population, soil, water and air pollution are unavoidable. The use
of living organisms to degrade pollutants in order to prevent pollution, restore natural surroundings
and prevent further pollution is known as bioremediation. This process employs microorganisms
such as bacteria, fungi, algae and others. Pollutants including pesticides are metabolized and used
as carbon or nitrogen sources for survival. Microorganisms can also be used to synthesize a number
of metabolites, including anabolites and catabolites. Recombinant DNA techniques can be used to
modify wild-type organisms.
Microorganisms can also be used to synthesize a range of metabolites, including pesticide
and other xenobiotic compounds. Traditional methods and recombinant DNA techniques are being
used to improve the productivity of wild-type organisms. This can be accomplished through the
selection of mutants with desired phenotypes in controlled environmental conditions. Microbial
techniques have attempted to compete with chemical product synthesis. Rising prices for petroleum,
a non-renewable resource, necessitates the use of microbial production. First, advances in metabolic
engineering techniques have resulted in engineered microorganisms that produce an excess of
metabolites and intermediates. Another significant advancement is “heterologous expression of
genes”, which involves introducing genes from other organisms into microbial cells, that then
express these foreign genes.
Second, the compounds produced by microorganisms are highly specific in terms of structure,
position and stereochemistry. As a result, the cost of the downstream process is reduced. Third,
unlike chemical synthesis, microbial processes do not use toxic heavy metals as catalysts, and the
operating temperatures and pressures are lower. Microorganisms can produce valuable chemicals as
well as degrade toxic pollutants. Sophisticated recombinant DNA techniques are used for targeted
genetic manipulations, changing existing pathways or heterologous expression of genes in a new
external pathway (Santos and Stephanopoulos 2008). However, the relationship between phenotype
and genotype in organisms is extremely complex. To achieve the desired phenotype, a combination
of genome editing tools are used.
The purpose of genetic improvement would be to develop non-growing but metabolically active
sedentary cells that would reroute the metabolic fluxes away from growth and towards product
formation. However, when it comes to products such as recombinant proteins that are intimately
related to the process of growth, it could be challenging to identify the genes that need to be
knocked-out or knocked-in in order to achieve the phenotype that is desired. In order to get around
the challenge, inverse metabolic engineering is the best option. Inverse metabolic engineering is the
process of elucidating a metabolic engineering technique through the identification, construction
or calculation of a desired phenotype, the identification of the genetic or specific environmental
elements conferring that the phenotype, and the intentional genetic or environmental alteration of
another strain or organism to give that phenotype. Inverse Metabolic Engineering (IME) is divided
into three steps:
i. Choosing or creating strains with the desired phenotype.
ii. Investigating the impact of genetic and environmental factors on the phenotype.
iii. Transferring the phenotype to a different organism.
An ideal balance between phenotypic expression and cell viability is essential since genes are
being overexpressed. The expression of several enzymes should be balanced (Koffas et al. 2003,
Pitera et al. 2007). All the enzymes involved in a pathway must work together. An imbalance
expression leads to accumulation of intermediate metabolites, potentially impairing cell viability.
This chapter will briefly discuss y how genetic modifications can be made in microbial cells
to achieve the desired phenotypic expression. Some prior knowledge about the pathways and the
metabolites is important prerequisite.
The applications of Pseudomonas spp. described in this chapter are divided into two parts.
To begin with, it entails the use of engineered microbes to degrade pollutants in the environment.
Pesticides of the class organophosphate and others are discussed here, including Benzene/
toluene/p-xylene (BTX), phenylurea, methyl parathion and cadmium; 1,2,3-trichloropropane;
phenol; chlorpyrifos; carbofuran; c-hexachlorocyclohexane (c-HCH) or lindane; S-triazines;
polycyclic aromatic hydrocarbons (PAHs); Diuron; Naphthol. Second, it includes ethylene,
polyhydroxyalkanoate, 2-methylcitric acid, isoprenoid, long-chain polyunsaturated fatty acids,
n-butanol, para-hydroxyl benzoic acid, and enzyme production using pollutants as a substrate
(Figure 8.1). Many conventional chemical synthesis methods rely on depleting petroleum as a
prerequisite. As a result, microorganisms are being used in the clean and green synthesis of a wide
range of petroleum products.
Pseudomonas putida
Parathion Ethylene
Value added products
Chlorpyrifos
Pesticides
Figure 8.1. Utilization of Pseudomonas putida for the degradation of pesticides and production of value-added
products, PAH-Polycyclic aromatic hydrocarbons, c-HCH-c Hexachlorocyclohexane, mcl-PHA, medium chain length
polyhydroxyalkanoates, short chain length polyhydroxyalkanoates, FDCA-2,5 furandicarboxyalic acid, Lc-PUFA-long chain
polyunsaturated fatty acid, PHBA-para-Hydroxy benzoic acid.
Pseudomonas putida: An Environment Friendly Bacterium 127
Microorganisms found in nature may have some restriction on degradation of substrate and
synthesis of a product. Such microorganism especially bacteria can be improved by making a
change in the genes using different tools available today. Furthermore, utilization of constitutive
promoter libraries can modulate gene expression. They have a wide range of expression levels and
can control multiple genes at the same time (Alper et al. 2005b, Jensen and Hammer 1998). This
kind of system also eliminates the need for inducible systems. It is extremely useful in large-scale
product production. Secondary structures of mRNA, RNase cleavage sites and RBS sequestering
sequences are the post-transcription modulators and control the expression of multiple genes under
the control of single promotor. Tunable Intergenic Regions (TIGR) placed between coding regions
aid in post-transcriptional manipulation of phenotype expression (Pfleger et al. 2006).
2022). TALENs and ZFNs can recognize one and three base pairs with modular regulatory proteins,
respectively. CRISPR (Clustered Regularly Interspaced Short Palindromic Repeats)-Cas is an
adaptive immune system that protects bacteria from foreign nucleic acids. CRISPR-Cas systems
are a type of adaptive immune defence mechanism found in prokaryotes. This mechanism can be
used to create specific double-stranded DNA breaks. NHEJ or HDR oversee repairing these sites
(Sharma and Shukla 2022). CRISPR has several benefits, including high precision, convenience of
design, fast turnaround time and is inexpensive. The disadvantage of CRISPR is that it may cause
cell death in bacteria that lack the NHEJ repair system. To combat this, recombinase proteins such as
SSr and -Red are used (Caldwell and Bell 2019). For example, 2,5--furandicarboxylic acid (FDCA)
is a renewable alternative to petroleum-based terephthalic acid and is used to make polyurethanes,
polyamides and polyesters (Pham et al. 2020). As a result, FDCA is a critical component of industrial
production. A number of enzymes convert 5-hydroxymethyl furfural (HMF) to FDCA (Caldwell
and Bell 2019, Pham et al. 2020 Zhang et al. 2002, Biot-Pelletier and Martin 2014). Recombinant
P. putida S12 expressing HMF/furfural oxireductase can produce FDCA where CRISPR-Cas was
used to induce double-stranded breaks. Double-stranded breaks were then used to recombine
HMF/furfural oxidoreductase through Red-mediated recombineering. The recombinant strain was
extremely efficient, converting approximately 88–85% of HMF to FDCA and frequent one-step
stable gene integration was possible into all chromosomes of polyploid P. putida S12. As a result,
the most stable strain of P. putida S12 was produced for the synthesis of FDCA. Although the
stable integration of genes and efficient degradation was observed in this recombinant strain, higher
concentration HMF (100 mM) was affecting the conversion. Initial increased density of biomass
rescues the inhibition with increased conversion 100 mM and 150 mM HMF to 96 to 75% of FDCA
(Hsu et al. 2020).
Genome shuffling is another method usually used for strain improvement at the industrial level.
It is a combination of DNA shuffling and directed evolution used to improve strains (Zhang et al.
2002). DNA shuffling has successfully achieved directed molecular evolution of several genes and
pathways (Biot-Pelletier and Martin 2014). For genome shuffling, the protoplast fusion method is
employed where the fusion of two cells with different genetic traits achieves the desired modified
phenotype. Even intergenic hybridization between Aspergillus niger and Penicillium digitatum
results in increased verbenol production (Rao et al. 2003). The consequence of protoplast fusion is
a library of strains with accumulated mutations. These strains can now be screened for phenotypes
of interest. As a result, it is a very useful tool, and more hybrid strains can be obtained than through
protoplast fusion as more parents are involved and saves a lot of assays and time (Zhang et al. 2002).
Pseudomonas are the gram-negative saprophytic in most of the soil dwelling bacteria. One
of the most important features of Pseudomonas sps., is degradation of almost all type of carbon
skeletons. The Recombinant DNA Advisory Committee has approved Pseudomonas putida as one
of the popular bacteria for genetic engineering and certified it as a biosafe host for heterologous gene
expression (Nelson et al. 2002). They exhibit rapid and robust growth. P. putida is resistant to very
high concentrations of toxic especially xenobiotic compounds due to solvent resistance mechanisms
mediated by the interaction of cellular factors (Ramos et al. 2002). Some examples include fine-
tuning lipid fluidity to adjust membrane functions, activating stress response systems and inducing
efflux pumps. P. putida is also metabolically extremely adaptable. It has a ED/EMP cycle, a distinct
metabolic pathway that employs enzymes from the Entner-Doudoroff (ED), pentose phosphate
and Embden-Meyerhof-Parnas (EMP) pathways (Nikel et al. 2015). Due to the availability of the
ED/EMP cycle, it is now possible to build completely different biochemistries using new
methodologies. In this chapter, the use of genetically modified Pseudomonas sp. especially P. putida
for compound synthesis and degradation to aid in pollution control will be described.
Pseudomonas putida: An Environment Friendly Bacterium 129
all three chemicals in a combination of the tod and tol metabolic pathways (Worsey and Williams
1975, Zylstra et al. 1988).
As a result of the inability of catechol 2,3-dioxygenase in the tod pathway to attack
3,6-dimethylpyrocatechol produced from p-xylene and xylene oxygenase in the athway to employ
benzene as a substrate, it is possible that complete biodegradation of the BTX combination through
these two pathways will not be possible (Worsey and Williams 1973, Zylstra et al. 1988, Gibson
et al. 1970, Gibson et al. 1974). Consequently, an existing route must be restructured. In order to
manipulate the genes for two pathways in P. putida, a hybrid metabolic pathway was constructed
around the crucial metabolic step. In P. putida, an attempt was made to mineralize BTX by a
hybrid route. To construct the hybrid pathway, a bridging step between the tod and tol pathways
was identified. The tol pathway destroys cis-glycol chemicals, which are metabolic intermediates
immediately before the catechol molecules in the tod pathway. As the only source of carbon, BCG
was 90% degraded (Lee et al. 1994). P. putida TB101 was generated by introducing the TOL plasmid
pWW0 into P. putida F39/D, a P. putida FI variant incapable of converting cis-glycol chemicals to
catechols. Toluate-cis-glycol dehydrogenase in P. putida TB101 diverted the metabolic flux of BTX
into the tod pathway at the level of cis-glycol molecules, resulting in the simultaneous mineralization
of BTX mixture without accumulation of any metabolic intermediates (Lee et al. 1994). Toluate-cis
glycol dehydrogenase, which is encoded on the TOL plasmid of P. putida mt-2, was considered to
be responsible for the degradation of these cis-glycol compounds due to the comparable chemical
structure of its original substrates, benzoate-cis-glycol (BACG) and TACG. BCG was utilized as
a substrate during the toluate-cis-glycol dehydrogenase experiment. The observation shows that
the reaction product produced by BCG’s toluate-cis-glycol dehydrogenase may be destroyed by
catechol 2,3-dioxygenase, the enzyme that follows toluate-cis-glycol dehydrogenase in the tol
pathway. In P. putida TBlOl, a hybrid pathway was developed for the complete mineralization
of benzene, toluene and p-xylene. Using P. putida TBlOl, simultaneous biodegradation of BTX
combinations was accomplished with maximum specific degradation rates of 0.27, 0.86 and
2.89 mgmg–1 biomass/h for benzene, toluene, and p-xylene, respectively (Lee et al. 1994).
8.3.3 Phenylurea
Phenylureas like diuron linuron isoproturon and chlorotoluron were once used to control weeds, but
they have now contaminated drinking water. Diuron is a phenylurea herbicide that is used to control
weeds in a number of crops (Moretto et al. 2019). It has been found in soil and water (Giacomazzi
and Cochet 2004). Diuron has been linked to kidney disease, haematopoiesis and hemolytic anaemia
(Ihlaseh-Catalano et al. 2014). In high salinity conditions, the CASB3 strain of Stenotrophomonas
rhizophila was used to degrade diuron. Fifty mg L–1 diuron was completely degraded in 48 to
120 hr (Silambarasan et al. 2020). Arthrobacter globiformis D47 can ‘partially’ degrade herbicide
diuron via urea carbonyl group hydrolysis (Cullington and Walker 1999, Turnbull et al. 2001). The
concentration of 3,4-dichloroaniline (DCA) rises as the diuron degrades. A very small amount of
DCA can be converted to 3,3’ 4,4’ tetrachloroazobenzene and the remaining bulk of the DCA continue
bound to the organic matter with slow half-life for several years. Pseudomonas putida strain was
enriched for the demineralization of DCA. The pathway investigation study suggested biodegradion
through 3,4 dichloromuconate, 3 chlorobutenolide, 3 chlorolevulinic acid, 3 chloromelyelacetate
3chloro 4 ketadipate to succinate (You and Bartha 1982).
(PNP) and dimethyl phosphorothioate but are unable to degrade PNP further. These bacteria were
used to clone ophc2, an organophosphorus hydrolase encoding gene. Cd can be immobilized
by bacteria such as P. aeruginosa KUCd1 and Pseudomonas sp. strain RB (Zhang et al. 2016).
A soil-dwelling bacterium was isolated having the potential of methyl parathion degradation ability.
The bacteria were responsible for producing the enzyme organophosphorus acid anhydrase, which
was responsible for the hydrolysis of methyl parathion into p-nitrophenol. Hydroquinone and 1, 2,
4-benzenetriol were produced eventually as a by-product of further degradation of p-nitrophenol.
At the end Maleyl acetate was produced by cleaving the last ring component, 1, 2, 4-benzenetriol,
which was catalyzed by the enzyme benzenetriol oxygenase (Rani and Lalithakumari 1994).
Pseudomonas sp., isolated from a mixed population of methyl parathion and parathion-degrading
bacteria, can hydrolyze the MP to p-nitrophenol and demands glucose and other carbon sources
for growth (Chaudhry et al. 1988). Flavobacterium sp. isolate from the same mixed population
could convert pnitrophenol to nitrite. It can take up to 48 hr for both bacteria to mineralize methyl
parathion as their only carbon source. Surface expression of metal-binding proteins in host strains is
an effective way to increase heavy metal immobilization. To maximize bioremediation, an EGFP-
expressing X3 bacterium was metabolically engineered to have MP degrading genes as well as a
Cd-immobilization phenotype (Zhang et al. 2016).
8.3.6 Phenol
Phenol is widely used and is a unfavourably industrial waste produced by petroleum refineries
(NPCS Board of Consultants and Engineers 2000). As it quickly penetrates to the skin and can cause
irritation to the eyes, respiratory tract and can be fatal in chronic exposure. The use is unavoidable
leaving an option of maintaining the permissive limit. The maximum allowed level is 2ppm and
can be maintained below or neutralized to zero level using P. putida. P. putida FNCC-0071 cells
immobilized in alginate beads were used to degrade phenol (Hannaford and Kuek 1999). About
600 ppm of phenol was degraded within 48 hr without affecting viability. Recently, a constant
132 Bioremediation for Sustainable Environmental Cleanup
electric current was applied for the enhancement of phenol degradation. Basically P. putida and
other microorganisms degrade via meta and para cleavage pathways influenced by the concentration,
temperature and biogenic factors. Most of the redox biochemical reaction are facilitated by applying
constant voltage across the bioreactor along with microbial community (Bandhyopadhyay et al.
1999).
8.3.8 Beta-cypermethrin
Beta-cypermethrin (β-CY) is a significant pyrethroid pesticide (Xiao et al. 2015). It is a toxin that
affects the reproductive, immune, nervous and genetic systems. (Jin et al. 2011, McKinlay et al.
2008, Schettgen et al. 2002). The pesticide can be co-metabolized and degraded by Aspergillus
niger YAT, Bacillus licheniformis B-1, Brevibacterium aureum and Brevibacillus parabrevis
BCP-09 (Tran et al. 2013, Chen et al. 2013, Deng et al. 2015, Tang et al. 2018). However, soil
degradation was limited. As a result, Loh and Wang (1997) discovered that P. putida ATCC 49451
can degrade 4-chlorophenol. The process was accelerated in the presence of sodium glutamate. But
this process was hampered by glucose (Zhao et al. 2019a,b).
Lindane or c-hexachlorocyclohexane, c-HCH is a highly toxic pesticide that cause Anabaena
to lose protective chaperons, inhibit photosynthesis and fix nitrogen (Bueno et al. 2004, Singh
1973, Nawab et al. 2003). In Sphingomonas paucimobilis B90, the linA2 gene encodes c-HCH
dehydrochlorinase, which degrades the pesticide. Similarly, the pesticide is degraded by the linA
gene in P. paucimobilis UT26 (Adhya et al. 1996, Nagata et al. 1997). In the presence of an inducible
promoter PpsbA1, these genes were knocked in and overexpressed in Anabaena. In 12 hr, this strain
could degrade 10 ppm of lindane isomers (Chaurasia et al. 2013).
8.3.9 s-Triazines
s-Triazines are a class of herbicides used to control weeds, such as atrazine. Pseudomonas sp., strain
A uses ozone to oxidize atrazine, and the ozonated products are used by soil microbes (Karns et
al. 1987, Kearney et al. 1988). S-triazine is also used as the only nitrogen source by Pseudomonas
(A, D and F) and Klebsiella pneumoniae strain 90 and 99 (Cook and Hutter 1981). Aminohydrolases,
which catalyze ring deamination, and amidohydrolases catalyze this conversion. These genes were
discovered through transposon mutagenesis (Eaton and Karns 1991). Recently pure culture of
Pseudomonas sp., ADP was identified which can mineralize the complete ring of atrazine to CO2
and hydroxyatrazine and polar metabolites.
contaminated soil (Kumar et al. 2012). After nitosoguanidine treatment, phenotypically favourable
strains were chosen for protoplast fusion (Dai et al. 2005, Singleton et al. 2009, Dai and Copley
2004, Gong et al. 2009). As a result, the SF-IOC11-16A strain created by genome shuffling was
efficient in degradation of PAHs (Kumar et al. 2012).
8.3.11 Naphthalene
Naphthalene degradation is associated with Pseudomonas putida PpG7 (Dunn and Gunsalus 1973),
P. paucimobilis Q1 (Kuhm et al. 1991) and P. putida ND69 (Song et al. 2018). In 48 hr, the latter
can degrade up to 98% of 2mg L–1 naphthalene. Jia Yan et al. (2008) observed the degradation of
naphthalene in Pseudomonas N7 in 2008; the degradation rate reached 95.66%. The naphthalene
degrading gene in P. putida ND6 is found on pND 6-1, a 102 kb plasmid. pND 6-1 has a G + C content
of 57%, which includes the oriV, the region related to plasmid replication and stable inheritance, and
the naphthalene degradation gene region. For naphthalene degradation, there are 23 coding domains
(Li et al. 2004). The genes involved are nahG, which codes for salicylic acid hydroxylase, and nahR,
that codes for NahR, a transcriptional regulatory protein of the LysR family. The gene nahY encodes
a naphthalene chemotactic protein, while the gene nahV encodes a salicylaldehyde dehydrogenase
(Song et al. 2018).
8.3.12 Toluene
The strain P. putida DOT-T1E is rifampin resistant. It can withstand supersaturating toluene
concentrations (Ramos-Gonza’lez et al. 2003). Toluene is used as a carbon and energy source (Ramos
et al. 1995). The mechanism is the tod-pathway, which converts toluene into 3-methylcathecol, is
then used in the Krebs cycle (Mosqueda et al. 1999). P. mendocina KR1 metabolizes toluene to
p-cresol via the T4MO pathway using the tmo gene product (Whited and Gibson 1991). Pcu genes
further oxidize this to 4-hydroxybenzoate. 4-HBA is a precursor to paraben and methylparaben,
which are used in the production of liquid glass and antimicrobial agents (Huang et al. 1992, Soni
et al. 2002). Toluene is also converted into 4-HBA by P. putida DOT-T1E. To prevent toluene and
4-HBA from being misrouted, the tod and beta-ketoadipate pathways were turned off to maximize
product formation (Ramos-Gonza’lez et al. 2003).
a high copy number and a broad host range that carried efe from P. syringae. The recombinant’s
maximum ethylene production rate was 26 times greater than that of the parental P. syringae. The
oxygen and carbon sources could be improved to increase yield even further (Wang et al. 2010).
As the chemical synthesis of 2-MC requires hazardous solvents such as benzene, microorganisms
are used. 2-MC was produced using Ralstonia eutropha H16 and P. putida KT2440 (Brämer and
Steinbüchel 2001). The manipulated strains, R. eutropha DeltaacnM (Re) OmegaKmprpC (Pp)
and P. putida DeltaacnM (Pp) OmegaKmprpC (Re), were created by inserting the 2-methylcitrate
synthase gene, causing the 2-methyl-cis-aconitate hydratase to be disrupted (acnM). Due to the
disruption of the acnM, there was an excessive generation of 2MC, which led to its build up. The
maximum concentrations attained by the stains after 144 hr of cultivation were 7.2 g L–1, which
is equivalent to 26.5 mM, and 19.2 g L–1, that is equivalent to 70.5 mM, respectively (Ewering
et al. 2006). Guaiacol is one of the products of depolymerization of kraft lignin along with catechol
benzoate and toluene. In recent years, more attention was diverted to convert guaiacol to a value
added product like muconic acid. P. putida KT2440 is engineered for two step the conversion of
guaiacol to muconic acid. Deletion of CatBC gene thereby blocking the catabolism to muconic and
insertion of cytochrome P450 and ferredoxin reductase gene from R. rhodochrous enabling the
conversion of guaiacol to catechol (Almqvist et al. 2021). Multiple natural enzymes in P. putida
KT2440 are capable of utilizing vanillin as a substrate. Vanillin dehydrogenase and various aldehyde
reductases are enzymes involved in the breakdown of vanillin into vanillyl alcohol and vanillic acid,
respectively (Simon et al. 2014). GN442PP 2426, which was previously modified to manufacture
vanillin from ferulic acid (Graf and Altenbuchner 2014, García-Hidalgo et al. 2020), may be a
more ideal host strain than KT2440 for the generation of VA since vanillic acid can then be further
assimilated via protocatechuate. This is because ferulic acid is converted into vanillin by genetically
engineering KT2440. P. putida EM42, a genome-reduced variety of P. putida KT2440 with superior
physiological features, was recently modified for growth on cellobiose (Dvořák). Cellobiose and
glucose can be used together in the same metabolic pathway owing to a mutant (PP_1444) that
lacks the periplasmic glucose dehydrogenase Gcd, but unfortunately the Δgcd strain suffered from
a significant growth defect. The growth defect was compensated by introduction of heterologous
glucose (Glf from Zymomonas mobilis) and cellobiose (LacY from Escherichia coli) transporters
with surprised production of pyruvate (Bujdoš et al. 2023)
8.4.5 Isoprenoid
It is a profitable molecule that has implications in the pharmaceutical, as well as the food and
beverage industries (Arendt et al. 2016). Bacteria such as P. putida can tolerate larger amounts
of isoprenoids (Mi et al. 2014). As a result, they can be utilized to satisfy an increasing demand.
P. putida is utilized in the process of biotransformation of isoprenoids in order to get oxidation
products of the plant monoterpene, limonene (Loeschcke and Thies 2015) or de novo biosynthesis
of the monoterpene geranic acid (Mi et al. 2014) or the carotenoids zeaxanthin and -carotene.
P. putida utilizes the methylerythritol 4-phosphate (MEP) pathway, whereas other bacteria utilize
the unrelated mevalonate (MVA) process to produce acetyl-CoA. P. putida KT2440 was genetically
modified to manufacture modest quantities of lycopene via the MEP route under the control of
IPTG-induced stress regulated promoters. These promoters allowed to produce measurable levels
of lycopene. The amount of lycopene that was produced by this strain increased by a factor of 50
(Hernandez-Arranz et al. 2019).
8.4.7 n-Butanol
Currently, n-butanol is produced from petroleum compounds. It is used as a precursor in the
production of polymers and paints (Cuenca et al. 2016). Microorganisms can also produce butanol
from municipal solid waste and lignocellulose materials (Ezeji et al. 2007). Butanol can be produced
at concentrations of 5mg L–1 in P. putida S12 by the heterologous expression of the Clostridium
acetobutylicum pathway (Nielsen et al. 2009). The gene glcB encodes an enzyme in the glyoxylate
shunt. The glcB mini-Tn5 mutant is mutated to prevent butanol assimilation.
the rice rhizospheric soil and mutated for improved production of laccase enzyme using UV and
chemical mutation. The mutant strain E4 was a hyper producer of laccase with 34.12 U.L–1 with
8.09 percent yield (Verma et al. 2016). Another strain of P. putida 922 was optimized to produce
lipase with 24 U ml yield of enzyme in 48 hr (Khan et al. 2014). Along with chemical fertilizers
P. putida have improved the production of soil enzyme and thereby improving crop yield (Nosheen
et al. 2018).
8.5 Conclusion
The growing population will only continue to grow adding a large amount of waste that may not
be degraded by natural degraders. As a result, the environment is in desperate need of remediation.
Remediation can be accomplished in a number of ways. Using biological organisms or products
to do the same is the most cost effective and least harmful to the environment. This is referred to
as bioremediation. This chapter describes how P. putida can be used as a bioremediation tool. It
is a very robust soil organism that is relatively easy to genetically modify. P. putida is rod shape,
gram-negative motile bacteria, ubiquitous in the environment, especially in pollutant-contaminated
water and soil. It thrives on almost all the carbon skeletons including plastic (Vague et al. 2019).
This organism is naturally equipped with various kinds of enzymes making them highly important
bacteria from the environmental point of view. Unlike E. coli, strains of P. putida are already adapted
to the soil environment. They can grow and degrade pollutants in the soil or groundwater more
easily. P. putida KT2440 has been bred to grow in anoxic conditions. As NADH produced during
metabolism must be re-oxidized, oxygen is required. This strain has also been genetically modified
to degrade pesticides (methyl parathion, chlorpyrifos, fenpropathrin, cypermethrin, carbofuran and
carbaryl) (Gong et al. 2018). Pseudomonas strains that are resistant to extreme temperatures, pH and
pressure can be created through further modification. As a result, the scope of bioremediation can
be expanded to include harsher conditions. Pseudomonas has a wide range of applications due to its
unique metabolism. It has demonstrated applications in both bioremediation and the production of
compounds that will aid in the reduction of reliance on petroleum products.
References
Adhya, T. K., S. K. Apte, K. Raghu, N. Sethunathan and N. B. Murthy. 1996. Novel polypeptides induced by the
insecticide lindane (gamma hexachlorocyclohexane) are required for its biodegradation by a Sphingomonas
paucimobilis strain. Biochem. Biophys. Res. Commun. 221(3): 755–61.
Adrio, J. L. and A. L. Demain. 2006. Genetic improvement of processes yielding microbial products. FEMS Microbiol.
Rev. 30(2): 187–214.
Allen, E. E. and D. H. Bartlett. 2002. Structure and regulation of the omega-3 polyunsaturated fatty acid synthase
genes from the deep-sea bacterium Photobacterium profundum strain SS9The GenBank accession numbers for
the sequences reported in this paper are AF409100 and AF467805. Microbiol. 148(6): 1903–1913.
Almqvist, H., H. Veras, K. Li, J. Garcia Hidalgo, C. Hulteberg, M. Gorwa-Grauslund, N. Skorupa Parachin and
M. Carlquist. 2021. Muconic acid production using engineered Pseudomonas putida KT2440 and a guaiacol
rich fraction derived from kraft lignin. ACS Sustain. Chem. Eng. 9(24): 8097–8106.
Alper, H., C. Fischer, E. Nevoigt and G. Stephanopoulos. 2005a. Tuning genetic control through promoter engineering.
Proc. Natl. Acad. Sci. USA. 102(36): 12678–12683.
Alper, H., Y. S. Jin, J. F. Moxley and G. Stephanopoulos. 2005b. Identifying gene targets for the metabolic engineering
of lycopene biosynthesis in Escherichia coli. Metab. Eng. 7(3): 155–164.
Alshammari, A., V. N. Kalevaru, A. Bagabas and A. Martin. 2016. Production of ethylene and its commercial
importance in the global market. pp. 82–115. In: H. Al-Megren (Ed.). Petrochemical Catalyst Materials
Processes And Emerging Technologies. IGI Global.
Amiri-Jami, M., H. Wang, Y. Kakuda and M. W. Griffiths. 2006. Enhancement of polyunsaturated fatty acid production
by Tn5 transposon in Shewanella baltica. Biotechnol. Lett. 28(15): 1187–1192.
Aoyama, M., K. Agari, G. H. Sun-Wada, M. Futai and Y. Wada. 2005. Simple and straightforward construction of a
mouse gene targeting vector using in vitro transposition reactions. Nucleic Acids Res. 33(5): e52–e52.
Arendt, P., J. Pollier, N. Callewaert and A. Goossens. 2016. Synthetic biology for production of natural and new‐to‐
nature terpenoids in photosynthetic organisms. Plant J. 87(1): 16–37.
138 Bioremediation for Sustainable Environmental Cleanup
Bandhyopadhyay, K., D. Das and B. R. Maiti. 1999. Solid matrix characterization of immobilized Pseudomonas
putida MTCC 1194 used for phenol degradation. Appl. Microbiol. Biotechnol. 51(6): 891–895.
Bang, S. W. 1997. Molecular analysis of p-nitrophenol degradation by Pseudomonas sp. strain ENV2030. Ph.D.
Thesis, Department of Environmental Sciences, Rutgers University, New Brunswick, New Jersey.
Barker, J. L. and J. W. Frost. 2001. Microbial synthesis of p‐hydroxybenzoic acid from glucose. Biotechnol. Bioeng.
76(4): 376–390.
Bhushan, B., A. Chauhan, S. K. Samanta and R. K. Jain. 2000. Kinetics of biodegradation of p-nitrophenol by
different bacteria. Biochem. Biophys. Res. Commun. 274(3): 626–630.
Bigley, A. N. and F. M. Raushel. 2013. Catalytic mechanisms for phosphotriesterases. Biochim.
Biophys. Acta Proteins Proteom. 1834(1): 443–453.
Biot-Pelletier, D. and V. J. Martin. 2014. Evolutionary engineering by genome shuffling. Appl. Microbiol. Biotechnol.
98(9): 3877–3887.
Bojanovič, K., I. D’Arrigo and K. S. Long. 2017. Global transcriptional responses to osmotic. Appl. Environ.
Microbiol. 83(7): e03236–16.
Brämer, C. O. and A. Steinbüchel. 2001. The methylcitric acid pathway in Ralstonia eutropha: new genes identified
involved in propionate metabolism The GenBank accession numbers for the nucleotide sequences of the prp
gene cluster are AF325554 and AF331923. Microbiol. 147(8): 2203–2214.
Bueno, M., M. F. Fillat, R. J. Strasser, R. Maldonado-Rodriguez, N. Marina, H. Smienk, C. Gómez-Moreno and
F. Barja. 2004. Effects of lindane on the photosynthetic apparatus of the Cyanobacterium anabaena. Environ.
Sci. Pollut. Res. Int. 11(2): 98–106.
Bujdoš, D., B. Popelářová, D. C. Volke, P. I. Nikel, N. Sonnenschein and P. Dvořák. 2023. Engineering of
Pseudomonas putida for accelerated co-utilization of glucose and cellobiose yields aerobic overproduction of
pyruvate explained by an upgraded metabolic model. Metabolic Eng. 75: 29–46.
Caldwell, B. J. and C. E. Bell. 2019. Structure and mechanism of the Red recombination system of bacteriophage λ.
Prog. Biophys. Mol. Biol. 147: 33–46.
Carvalho, F. D., I. Machado, M. S. Martínez, A. Soares and L. Guilhermino. 2003. Use of atropine-treated Daphnia
magna survival for detection of environmental contamination by acetylcholinesterase inhibitors. Ecotoxicol
Environ Saf. 54(1): 43–46.
Cha, D., H. S. Ha and S. K. Lee. 2020. Metabolic engineering of Pseudomonas putida for the production of various
types of short-chain-length polyhydroxyalkanoates from levulinic acid. Bioresour. Technol. 309: 123332.
Chagué, V., Y. Elad, R. Barakat, P. Tudzynski and A. Sharon. 2002. Ethylene biosynthesis in Botrytis cinerea. FEMS
Microbiol. Ecol. 40(2): 143–149.
Chaudhry, G. R., A. N. Ali and W. B. Wheeler. 1988. Isolation of a methyl parathion-degrading Pseudomonas sp.
that possesses DNA homologous to the opd gene from a Flavobacterium sp. Appl. Environ. Microbiol. 54(2):
288–93.
Chaurasia, A. K., T. K. Adhya and S. K. Apte. 2013. Engineering bacteria for bioremediation of persistent
organochlorine pesticide lindane (γ-hexachlorocyclohexane). Bioresour. Technol. 149: 439–445.
Chavarría, M., P. I. Nikel, D. Pérez-Pantoja and V. de Lorenzo. 2013. The Entner–doudoroff pathway empowers
Pseudomonas putida KT 2440 with a high tolerance to oxidative stress. Environ. Microbiol. 15(6): 1772–1785.
Chen, S., Y. H. Dong, C. Chang, Y. Deng, X. F. Zhang, G. Zhong, H. Song, M. Hu and L. H. Zhang. 2013.
Characterization of a novel cyfluthrin-degrading bacterial strain Brevibacterium aureum and its biochemical
degradation pathway. Bioresour. Technol. 132: 16–23.
Cook, A. M. and R. Huetter. 1981. s-Triazines as nitrogen sources for bacteria. J. Agric. Food Chem. 29(6): 1135–
1143.
Cuenca, M. D. S., C. Molina-Santiago, M. R. Gómez-García and J. L. Ramos. 2016. A Pseudomonas putida double
mutant deficient in butanol assimilation: a promising step for engineering a biological biofuel production
platform. FEMS Microbiol. Lett. 363(5): fnw018.
Cullington, J. E. and A. Walker. 1999. Rapid biodegradation of diuron and other phenylurea herbicides by a soil
bacterium. Soil Biol. Biochem. 31(5): 677–686.
Dai, M. and S. D. Copley. 2004. Genome shuffling improves degradation of the anthropogenic pesticide
pentachlorophenol by Sphingobium chlorophenolicum ATCC 39723. Appl. Environ. Microbiol. 70(4): 2391–
2397.
DeFrank, J. J. 1991. Organophosphorus cholinesterase inhibitors: detoxification by microbial enzymes. Applications
of Enzyme Biotechnology. Springer, Boston, MA.
Dejonghe, W., J. Goris, S. El-Fantroussi, M. Höfte, P. DeVos, W. Verstraete and E. M. Top. 2000. Effect of
dissemination of 2,4-dichlorophenoxyacetic acid (2,4-D) degradation plasmids on 2,4-D degradation and on
bacterial community structure in two different soil horizons. Appl. Environ. Microbiol. 66(8): 3297–3304.
Pseudomonas putida: An Environment Friendly Bacterium 139
Deng, W., D. Lin, K. Yao, H. Yuan, Z. Wang, J. Li, L. Zou, X. Han, K. Zhou, L. He and X. Hu. 2015. Characterization
of a novel β-cypermethrin-degrading Aspergillus niger YAT strain and the biochemical degradation pathway
of β-cypermethrin. Appl. Microbiol. Biotechnol. 99(19): 8187–8198.
Dunn, N. W. and I. C. Gunsalus. 1973. Transmissible plasmid coding early enzymes of naphthalene oxidation in
Pseudomonas putida. J. Bacteriol. 114(3): 974–979.
Dvořák, P. and V. de Lorenzo. 2018. Refactoring the upper sugar metabolism of Pseudomonas putida for co-utilization
of cellobiose, xylose, and glucose. Metab. Eng. 48: 94–108.
Dvorak, P., S. Bidmanova, J. Damborsky and Z. Prokop. 2014. Immobilized synthetic pathway for biodegradation of
toxic recalcitrant pollutant 1,2,3-Trichloropropane. Environ. Sci. Technol. 48(12): 6859–6866.
Eaton, R. W. and J. S. Karns. 1991. Cloning and comparison of the DNA encoding ammelide aminohydrolase and
cyanuric acid amidohydrolase from three s-triazine-degrading bacterial strains. J. Bacteriol. 173(3): 1363–
1366.
Ebert, B. E., F. Kurth, M. Grund, L. M. Blank and A. Schmid. 2011. Response of Pseudomonas putida KT2440 to
increased NADH and ATP demand. Appl. Environ. Microbiol. 77(18): 6597–6605.
El-sayed, G. M., N. A. Abosereih, S. A. Ibrahim, A. El-Razik, B. Ashraf, M. A. Hammad and F. M. Hafez. 2019.
Cloning of the Organophosphorus Hydrolase (oph) gene and enhancement of Chlorpyrifos degradation in the
Achromobacter xylosoxidans Strain GH9OP via Mutation Induction. Jordan J. Biol. Sci. 12(3).
Ewering, C., F. Heuser, J. K. Benölken, C. O. Brämer and A. Steinbüchel. 2006. Metabolic engineering of strains of
Ralstonia eutropha and Pseudomonas putida for biotechnological production of 2-methylcitric acid. Metab.
Eng. 8(6): 587–602.
Ezeji, T., N. Qureshi and H. P. Blaschek. 2007. Butanol production from agricultural residues: impact of degradation
products on Clostridium beijerinckii growth and butanol fermentation. Biotechnol. Bioeng. 97(6): 1460–1469.
Gahlawat, G. and S. K. Soni. 2017. Valorization of waste glycerol for the production of poly (3-hydroxybutyrate)
and poly (3-hydroxybutyrate-co-3-hydroxyvalerate) copolymer by Cupriavidus necator and extraction in a
sustainable manner. Bioresour. Technol. 243: 492–501.
Garcia, R., D. Pistorius, M. Stadler and R. Müller. 2011. Fatty acid-related phylogeny of myxobacteria as an approach
to discover polyunsaturated omega-3/6 fatty acids. J. Bacteriol. 193(8): 1930–1942.
García-Hidalgo, J., D. P. Brink, K. Ravi, C. J. Paul, G. Lidénb and M. F. Gorwa-Grauslund. 2020. Vanillin production
in Pseudomonas: whole-genome sequencing of Pseudomonas sp. strain 9.1 and reannotation of Pseudomonas
putida CalA as a vanillin reductase. Appl. Environ. Microbiol. 86: e02442–19.
Gemperlein, K., G. Zipf, H. S. Bernauer, R. Müller and S. C. Wenzel. 2016. Metabolic engineering of Pseudomonas
putida for production of docosahexaenoic acid based on a myxobacterial PUFA synthase. Metab. Eng.
33: 98–108.
Gemperlein, K., S. Rachid, R. O. Garcia, S. C. Wenzel and R. Müller. 2014. Polyunsaturated fatty acid biosynthesis in
myxobacteria: different PUFA synthases and their product diversity. Chem. Sci. 5(5): 1733–1741.
Giacomazzi, S. and N. Cochet. 2004. Environmental impact of diuron transformation: a review. Chemosphere.
56(11): 1021–1032.
Gibson, D. T., G. E. Cardini, F. C. Maseles and R. E. Kallio. 1970. Oxidative degradation of aromatic hydrocarbons
by microorganisms. Biochem. 9(7): 1631–1635.
Gibson, D. T., V. Mahadevan and J. F. Davey. 1974. Bacterial metabolism of para-and meta-xylene: oxidation of the
aromatic ring. J. Bacteriol. 119(3): 930–936.
Gong, J., H. Zheng, Z. Wu, T. Chen and X. Zhao. 2009. Genome shuffling: progress and applications for phenotype
improvement. Biotechnol. Adv. 27(6): 996–1005.
Gong, T., R. Liu, Y. Che, X. Xu, F. Zhao, H. Yu, C. Song, Y. Liu and C. Yang. 2016. Engineering Pseudomonas putida
KT 2440 for simultaneous degradation of carbofuran and chlorpyrifos. Microb. Biotechnol. 9(6): 792–800.
Gong, T., X. Xu, Y. Che, R. Liu, W. Gao, F. Zhao, H. Yu, J. Liang, P. Xu, C. Song and C. Yang. 2017. Combinatorial
metabolic engineering of Pseudomonas putida KT2440 for efficient mineralization of 1,2,3-trichloropropane.
Sci. Rep. 7(1): 44896.
Gong, T., X. Xu, Y. Dang, A. Kong, Y. Wu, P. Liang, S. Wang, H. Yu, P. Xu and C. Yang. 2018. An engineered
Pseudomonas putida can simultaneously degrade organophosphates. Sci. Total Environ. 628: 1258–1265.
Gosset, G. 2009. Production of aromatic compounds in bacteria. Curr. Opin. Biotechnol. 20(6): 651–658.
Graf, N. and J. Altenbuchner. 2011. Development of a method for markerless gene deletion in Pseudomonas putida.
Appl. Environ. Microbiol. 77(15): 5549–5552.
Graf, N. and J. Altenbuchner. 2014. Genetic engineering of Pseudomonas putida KT2440 for rapid and high-yield
production of vanillin from ferulic acid. Appl. Microbiol. Biotechnol. 98: 137–149.
Gunsalus, I. C. and G. C. Wagner. 1978. Bacterial P-450cam methylene monooxygenase components: cytochrome m,
putidaredoxin, and putidaredoxin reductase. In: J. A. Tainer (Ed.). Methods in Enzymology. Academic Press.
140 Bioremediation for Sustainable Environmental Cleanup
Hannaford, A. M. and C. Kuek. 1999. Aerobic batch degradation of phenol using immobilized Pseudomonas putida.
J. Ind. Microbiol. Biotechnol. 22(2): 121–126.
Hernandez-Arranz, S., J. Perez-Gil, D. Marshall-Sabey and M. Rodriguez-Concepcion. 2019. Engineering
Pseudomonas putida for isoprenoid production by manipulating endogenous and shunt pathways supplying
precursors. Microb. Cell Fact. 18(1): 41640.
Hsu, C. T. and Y. C. Kuo, Y. C. Liu and S. L. Tsai. 2020. Green conversion of 5-hydroxymethylfurfural to furan-2,5
dicarboxylic acid by heterogeneous expression of 5-hydroxymethylfurfural oxidase in Pseudomonas putida
S12. Microb. Biotechnol. 13(4): 1094–1102.
Huang, K., Y. G. Lin and H. H. Winter. 1992. p-Hydroxybenzoate ethylene terephthalate copolyester: structure of
high-melting crystals formed during partially molten state annealing. Polymer. 33(21): 4533–4537.
Ibeh, C. C. 2011. Thermoplastic Materials: Properties, Manufacturing Methods, and Applications. CRC Press.
Ihlaseh-Catalano, S. M., K. A. Bailey, A. P. F. Cardoso, H. Ren, R. C. Fry and J. L. V. deCamargo and D. C. Wolf.
2014. Dose and temporal effects on gene expression profiles of urothelial cells from rats exposed to diuron.
Toxicology. 325: 21–30.
Janssen, D. B., J. E. Oppentocht and G. J. Poelarends. 2001. Microbial dehalogenation. Curr. Opin. Biotechnol. 12(3):
254–258.
Jensen, P. R. and K. Hammer. 1998. Artificial promoters for metabolic optimization. Biotechnol. Bioeng. 58(2‐3):
191–195.
Jia Yan, H. Yin, J. S. Ye, H. Peng, B. Y. He, H. M. Qin, N. Zhang and J. Qiang. 2008. Characteristics and pathway of
naphthalene degradation by Pseudomonas sp. N7. Huan Jing Ke Xue. 29(3): 756–762.
Jin, Y., S. Zheng and Z. Fu. 2011. Embryonic exposure to cypermethrin induces apoptosis and immunotoxicity in
zebrafish (Danio rerio). Fish Shellfsh Immunol. 30(4-5):1049–1054.
Jones, K. C., J. A. Stratford, K. S. Waterhouse, E. T. Furlong, W. Giger, R. A. Hites, C. Schaffner and A. E. Johnston.
1989. Increases in the polynuclear aromatic hydrocarbon content of an agricultural soil over the last century.
Environ. Sci. Technol. 23(1): 95–101.
Kamravamanesh, D., T. Kovacs, S. Pflügl, I. Druzhinina, P. Kroll, M. Lackner and C. Herwig. 2018. Increased
poly-β-hydroxybutyrate production from carbon dioxide in randomly mutated cells of cyanobacterial strain
Synechocystis sp. PCC 6714: mutant generation and characterization. Bioresour. Technol. 266: 34–44.
Karns, J. S., M. T. Muldoon, W. W. Mulbry, M. K. Derbyshire and P. C. Kearney. 1987. Use of microorganisms
and microbial systems in the degradation of pesticides. pp. 156–170. In: H. M. LeBaron, R. O. Mumma,
R. C. Honeycutt, J. H. Duesing, J. F. Phillips and M. J. Haas [Eds.]. Biotechnology in Agricultural Chemistry.
American Chemical Society.
Karns, J. S., W. W. Mulbry, J. O. Nelson and P. C. Kearney. 1986. Metabolism of carbofuran by a pure bacterial
culture. Pestic. Biochem. Physiol. 25(2): 211–217.
Karpf, M. and R. Trussardi. 2009. Efficient Access to oseltamivir phosphate (Tamiflu) via the O‐Trimesylate of
Shikimic Acid Ethyl Ester. Angew Chem. Int. Ed. Engl. 48(31): 5760–5762.
Kaulmann, U. and C. Hertweck. 2002. Biosynthesis of polyunsaturated fatty acids by polyketide synthases. Angew
Chem. Int. Ed. Engl. 41(11): 1866–1869.
Kearney, P. C., M. T. Muldoon, C. J. Somich, J. M. Ruth and D. J. Voaden. 1988. Biodegradation of ozonated atrazine
as a wastewater disposal system. J. Agric. Food Chem. 36(6): 1301–1306.
Khan, F. H. N., A. U. Rehman and Z. Hussain. 2014. Production and partial characterization of lipase from
Pseudomonas putida. Ferment. Technol. 2: 112–119.
Kim, T. K., Y. M. Jung, M. T. Vo, S. Shioya and Y. H. Lee. 2006. Metabolic engineering and characterization of
phaC1 and phaC2 genes from Pseudomonas putida KCTC1639 for overproduction of medium‐chain‐length
polyhydroxyalkanoate. Biotechnol. Prog. 22(6): 1541–1546.
Kim, Y., D. A. Webster and B. C. Stark. 2005. Improvement of bioremediation by Pseudomonas and Burkholderia
by mutants of the Vitreoscilla hemoglobin gene (vgb) integrated into their chromosomes. J. Ind. Microbiol.
Biotechnol. 32(4): 148–154.
Kniel, L., O. Winter and K. Stork 1980. Ethylene: Keystone to the Petrochemical Industry (ChemicalIndustries).
Marcel Dekker.
Koçar, G. and N. Civaş. 2013. An overview of biofuels from energy crops: current status and future prospects. Renew.
Sustain. Energy Rev. 28: 900–916.
Koffas, M. A., G. Y. Jung and G. Stephanopoulos. 2003. Engineering metabolism and product formation in
Corynebacterium glutamicum by coordinated gene overexpression. Metab. Eng. 5(1): 32–41.
Koller, M., P. Hesse, H. Fasl, F. Stelzer and G. Braunegg. 2017. Study on the effect of levulinic acid on whey-based
biosynthesis of poly (3-hydroxybutyrate-co-3-hydroxyvalerate) by Hydrogenophaga pseudoflava. Appl. Food
Biotechnol. 4(2): 65–78.
Pseudomonas putida: An Environment Friendly Bacterium 141
Koma, D., H. Yamanaka, K. Moriyoshi, T. Ohmoto and K. Sakai. 2012. Production of aromatic compounds by
metabolically engineered Escherichia coli with an expanded shikimate pathway. Appl. Environ. Microbiol.
78(17): 6203–6216.
Kondakova, T. and J. E. Cronan. 2019. Transcriptional regulation of fatty acid cis–trans isomerization in the solvent‐
tolerant soil bacterium. Environ. Microbiol. 21(5): 1659–1676.
Krömer, J. O., D. Nunez-Bernal, N. J. Averesch, J. Hampe, J. Varela and C. Varela. 2013. Production of aromatics in
Saccharomyces cerevisiae—a feasibility study. J. Biotechnol. 163(2): 184–193.
Kuhm, A. E., A. Stolz and H. J. Knackmuss. 1991. Metabolism of naphthalene by the biphenyl-degrading bacterium
Pseudomonas paucimobilis Q1. Biodegrad. 2(2): 115–120.
Kulakova, A. N., M. J. Larkin and L. A. Kulakov.1997. The plasmid-located haloalkane dehalogenase gene from
Rhodococcus rhodochrous NCIMB 13064. Microbiol. 143(1): 109–115.
Kumar, M., M. P. Singh and D. K. Tuli. 2012. Genome Shuffling of Pseudomonas sp. Ioca11 for improving degradation
of polycyclic aromatic hydrocarbons. Adv. Microbiol. 2: 26–30.
Lee, J. Y., J. R. Roh and H. S. Kim. 1994. Metabolic engineering of Pseudomonas putida for the simultaneous
biodegradation of benzene. Biotechnol. Bioeng. 43(11): 1146–1152.
Lemoigne, M. 1926. Products of dehydration and of polymerization of β-hydroxybutyric acid. Bull. Soc. Chem. Biol.
8: 770–782.
Lenihan-Geels, G., K. S. Bishop and L. R. Ferguson. 2013. Alternative sources of omega-3 fats: can we find a
sustainable substitute for fish? Nutrients. 5(4): 1301–1315.
Li, W., J. Shi, X. Wang, Y. Han, W. Tong, L. Ma, B. Liu and B. Cai. 2004. Complete nucleotide sequence and
organization of the naphthalene catabolic plasmid pND6-1 from Pseudomonas sp. Gene. 336(2): 231–240.
Liu, K. and S. Li. 2020. Biosynthesis of fatty acid-derived hydrocarbons: perspectives on enzymology and enzyme
engineering. Curr. Opin. Biotechnol. 62: 41821.
Loeschcke, A. and S. Thies. 2015. Pseudomonas putida—a versatile host for the production of natural products. Appl.
Microbiol. Biotechnol. 99(15): 6197–6214.
Loh, K. C. and S. J. Wang. 1997. Enhancement of biodegradation of phenol and a nongrowth substrate 4-chlorophenol
by medium augmentation with conventional carbon sources. Biodegrad. 8(5): 329–338.
Lorente-Cebrian, S., A. G. V. Costa, S. Navas-Carretero, M. Zabala, J. A. Martinez and M. J. Moreno-Aliaga. 2013.
Role of omega-3 fatty acids in obesity, metabolic syndrome, and cardiovascular diseases: a review of the
evidence. J. Physiol. Biochem. 69(3): 633–651.
Malhat, F. and I. Nasr. 2011. Organophosphorus pesticides residues in fish samples from the River Nile tributaries in
Egypt. Bull. Environ. Contam. Toxicol. 87(6): 689–692.
Mallick, K., K. Bharati, A. Banerji, N. A. Shakil and N. Sethunathan. 1999. Bacterial degradation of chlorpyrifos in
pure cultures and in soil. Bull. Environ. Contam. Toxicol. 62(1): 48–54.
Mandalakis, M., N. Panikov, S. Dai, S. Ray and B. L. Karger. 2013. Comparative proteomic analysis reveals
mechanistic insights into Pseudomonas putida F1 growth on benzoate and citrate. AMB Express. 3(1): 41275.
McKinlay, R., J. A. Plant, J. N. B. Bell and N. Voulvoulis. 2008. Endocrine disrupting pesticides: implications for risk
assessment. Environ. Int. 34(2): 168–183.
Metz, J. G., P. Roessler, D. Facciotti, C. Levering, F. Dittrich, M. Lassner, R. Valentine, K. Lardizabal, F. Domergue,
A. Yamada and K. Yazawa. 2001. Production of polyunsaturated fatty acids by polyketide synthases in both
prokaryotes and eukaryotes. Science 293(5528): 290–293.
Mi, J., D. Becher, P. Lubuta, S. Dany, K. Tusch, H. Schewe, M. Buchhaupt and J. Schrader. 2014. De novo
production of the monoterpenoid geranic acid by metabolically engineered Pseudomonas putida. Microb. Cell
Factories. 13(1): 44866.
Monga, D., P. Kaur and B. Singh. 2021. Microbe mediated remediation of dyes, explosive waste and polyaromatic
hydrocarbons, pesticides and pharmaceuticals. Curr. Res. Microb. Sci. 3: 100092.
Moretto, J. A. S., J. P. R. Furlan, A. F. T. Fernandes, A. Bauermeister, N. P. Lopes and E. G. Stehling. 2019. Alternative
biodegradation pathway of the herbicide diuron. Int. Biodeterior. Biodegrad. 143: 104716.
Mosqueda, G., M. I. Ramos-González and J. L. Ramos. 1999. Toluene metabolism by the solvent-tolerant
Pseudomonas putida DOT-T1 strain, and its role in solvent impermeabilization. Gene 232(1): 69–76.
Müller, R., A. Wagener, K. Schmidt and E. Leistner. 1995. Microbial production of specifically ring-13C-labelled
4-hydroxybenzoic acid. Appl. Microbiol. Biotechnol. 43(6): 985–988.
Nagata, Y. K. Miyauchi, J. Damborsky, K. Manova, A. Ansorgová and M. Takagi. 1997. Purification and
characterization of a haloalkane dehalogenase of a new substrate class from a gamma-hexachlorocyclohexane
degrading bacterium. Appl. Environ. Microbiol. 63(9): 3707-3710.
Napier, J. A. 2002. Plumbing the depths of PUFA biosynthesis: a novel polyketide synthase-like pathway from marine
organisms. Trends Plant Sci. 7(2): 51–54.
142 Bioremediation for Sustainable Environmental Cleanup
Nawab, A., A. Aleem and A. Malik. 2003. Determination of organochlorine pesicides in agricultural soil with special
reference to γ-HCH degradation by Pseudomonas strains. Bioresour. Technol. 88(1): 41–46.
Nelson, K. E., C. Weinel, I. T. Paulsen, R. J. Dodson, H. Hilbert, V. A. P. Martinsdos Santos, D. E. Fouts,
S. R. Gill, M. Pop, M. Holmes and L. Brinkac. 2002. Complete genome sequence and comparative analysis of
the metabolically versatile Pseudomonas putida KT2440. Environ. Microbiol. 4(12): 799–808.
Neumann, G., R. Teras, L. Monson, M. Kivisaar, F. Schauer and H. J. Heipieper. 2004. Simultaneous degradation
of atrazine and phenol by Pseudomonas sp. strain ADP: effects of toxicity and adaptation Appl. Environ.
Microbiol. 70(4): 1907–1912.
Ng, C. Y., I. Farasat, C. D. Maranas and H. M. Salis. 2015. Rational design of a synthetic Entner–Doudoroff pathway
for improved and controllable NADPH regeneration. Metab. Eng. 29: 86–96.
Nickerson, D. P. and L. L. Wong. 1997. The dimerization of Pseudomonas putida cytochrome P450cam: practical
consequences and engineering of a monomeric enzyme. Protein Eng. 10(12): 1357–1361.
Nielsen, D. R., E. Leonard, S. H. Yoon, H. C. Tseng, C. Yuan and K. L. J. Prather. 2009. Engineering alternative
butanol production platforms in heterologous bacteria. Metab. Eng. 11(4-5): 262–273.
Nikel, P. I. and V. de Lorenzo. 2013. Engineering an anaerobic metabolic regime in Pseudomonas putida KT2440 for
the anoxic biodegradation of 1, 3-dichloroprop-1-ene. Metab. Eng. 15: 98–112.
Nikel, P. I., M. Chavarría, T. Fuhrer, U. Sauer and V. DeLorenzo. 2015. Pseudomonas putida KT2440 Strain
metabolizes glucose through a cycle formed by enzymes of the Entner-Doudoroff, Embden-Meyerhof-Parnas,
and pentose phosphate pathways. J. Biol. Chem. 290(43): 25920–25932.
Nosheen, A., H. Yasmin, R. Naz, A. Bano, R. Keyani and I. Hussain. 2018. Pseudomonas putida improved soil
enzyme activity and growth of kasumbha under low input of mineral fertilizers. Soil Sci. Plant Nutr. 64(4):
520–525.
Novackova, I., D. Kucera, J. Porizka, I. Pernicova, P. Sedlacek, M. Koller, A. Kovalcik and S. Obruca. 2019. Adaptation
of Cupriavidus necator to levulinic acid for enhanced production of P (3HB-co-3HV) copolyesters. Biochem.
Eng. J. 151: 107350.
NPCS Board of Consultants and Engineers. 2000. Water and Air Effluents Treatment. pp. 317. India: Asia pacific
Press business press.
Pfleger, B. F., D. J. Pitera, C. D. Smolke and J. D. Keasling. 2006. Combinatorial engineering of intergenic regions in
operons tunes expression of multiple genes. Nat. Biotechnol. 24(8): 1027–1032.
Pham, N. N., C. Y. Chen, H. Li, M. T. T. Nguyen, P. K. P. Nguyen, S. L. Tsai, J. Y. Chou, T. C. Ramliand and
Y. C. Hu. 2020. Engineering Stable Pseudomonas putida S12 by CRISPR for 2,5-Furandicarboxylic Acid
(FDCA) Production. ACS Synth. Biol. 9(5): 1138–1149.
Pitera, D. J., C. J. Paddon, J. D. Newman and J. D. Keasling. 2007. Balancing a heterologous mevalonate pathway for
improved isoprenoid production in Escherichia coli. Metab. Eng. 9(2): 193–207.
Poblete-Castro, I., J. Becker, K. Dohnt, V. M. Dos-Santos and C. Wittmann. 2012. Industrial biotechnology of
Pseudomonas putida and related species. Appl. Microbiol. Biotechnol. 93(6): 2279–2290.
Ramanathan, M. P. and D. Lalithakumari. 1999. Complete mineralization of methylparathion by Pseudomonas sp.
A3. Appl. Biochem. Biotechnol. 80(1): 1–12.
Ramos, J. L., E. Duque, M. J. Huertas and A. L. I. HaÏDour. 1995. Isolation and expansion of the catabolic potential
of a Pseudomonas putida strain able to grow in the presence of high concentrations of aromatic hydrocarbons.
J. Bacteriol. 177(14): 3911–3916.
Ramos, J. L., E. Duque, M. T. Gallegos, P. Godoy, M. I. Ramos-Gonzalez, A. Rojas, W. Teránand and A. Segura.
2002. Mechanisms of solvent tolerance in gram-negative bacteria. Annual Rev. Microbiol. 56(1): 743–768.
Ramos-González, M. I., A. Ben-Bassat, M. J. Campos and J. L. Ramos. 2003. Genetic engineering of a highly solvent-
tolerant Pseudomonas putida strain for biotransformation of toluene to p-hydroxybenzoate. Appl. Environ.
Microbiol. 69(9): 5120–5127.
Rani, N. L. and D. Lalithakumari. 1994. Degradation of methyl parathion by Pseudomonas putida. Can. J.
Microbiol. 40(12): 1000–1006.
Rao, S. C., R. Rao and R. Agrawal. 2003. Enhanced production of verbenol, a highly valued food flavourant, by an
intergeneric fusant strain of Aspergillus niger and Penicillium digitatum. Biotechnol. Appl. Biochem. 37(2):
145–147.
Rehm, B. H. and A. Steinbüchel. 1999. Biochemical and genetic analysis of PHA synthases and other proteins
required for PHA synthesis. Int. J. Biol. Macromol. 25(1-3): 43525.
Reyrat, J. M., V. Pelicic, B. Gicquel and R. Rappuoli. 1998. Counter selectable markers: untapped tools for bacterial
genetics and pathogenesis. Infect Immun. 66(9): 4011–4017.
Rink, R., M. Fennema, M. Smids, U. Dehmel and D. B. Janssen. 1997. Primary structure and catalytic mechanism of
the epoxide hydrolase from Agrobacterium radiobacterAD1. J. Biol. Chem. 272(23): 14650–14657.
Pseudomonas putida: An Environment Friendly Bacterium 143
Santos, C. N. S. and G. Stephanopoulos. 2008. Combinatorial engineering of microbes for optimizing cellular
phenotype. Curr. Opin. Chem. Biol. 12(2): 168–176.
Sauer, U. 2001. Evolutionary engineering of industrially important microbial phenotypes. Metab. Eng. 73: 129–169.
Schettgen, T., U. Heudorf, H. Drexler and J. Angerer. 2002. Pyrethroid exposure of the general population-is this due
to diet. Toxicol. Lett. 134(1-3): 141–145.
Schmack, G., V. Gorenflo and A. Steinbüchel. 1998. Biotechnological production and characterization of polyesters
containing 4-hydroxyvaleric acid and medium-chain-length hydroxyalkanoic acids. Macromolecules. 31(3):
644–649.
Serdar, C. M., D. C. Murdock and M. F. Rohde. 1989. Parathion hydrolase gene from Pseudomonas diminuta MG:
Subcloning, complete nucleotide sequence, and expression of the mature portion of the enzyme in Escherichia
coli. Nat. Biotechnol. 7(11): 1151–1155.
Sharma, B. and P. Shukla. 2022. Futuristic avenues of metabolic engineering techniques in bioremediation. Biotechnol.
Appl. Biochem. 69(1): 51–60.
Shelton, D. R. and C. J. Somich. 1988. Isolation and characterization of coumaphos-metabolizing bacteria from cattle
dip. Appl. Environ. Microbiol. 54(10): 2566–2571.
Shen, Y. J., P. Lu, H. Mei, H. J. Yu, Q. Hong and S. P. Li. 2010. Isolation of a methyl parathion-degrading strain
Stenotrophomonas sp. SMSP-1 and cloning of the ophc2 gene. Biodegrad. 21(5): 785–792.
Silambarasan, S., P. Logeswari, A. Ruiz, P. Cornejo and V. R. Kannan. 2020. Influence of plant beneficial
Stenotrophomonas rhizophila strain CASB3 on the degradation of diuron-contaminated saline soil and
improvement of Lactuca sativa growth. Environ. Sci. Pollut. Res. Int. 27(28): 35195–35207.
Simon, O., I. Klaiber, A. Huber and J. Pfannstiel. 2014. Comprehensive proteome analysis of the response of
Pseudomonas putida KT2440 to the flavor compound vanillin. J. Proteomics. 109: 212–227.
Singh, P. K. 1973. Effect of pesticides on blue-green algae. Arch. Mikrobiol. 89(4): 317–320.
Singleton, D. R., L. Guzmán-Ramirez and M. D. Aitken. 2009. Characterization of a polycyclic aromatic hydrocarbon
degradation gene cluster in a phenanthrene-degrading Acidovorax strain. Appl. Environ. Microbiol. 75(9):
2613–2620.
Song, F., Y. Shi, S. Jia, Z. Tan and H. Zhao. 2018. Advances of naphthalene degradation in Pseudomonas putida ND6.
Front. Bioeng. Biotechnol. 944(1): 20074.
Soni, M. G., S. L. Taylor, N. A. Greenberg and G. A. Burdock. 2002. Evaluation of the health aspects of methyl
paraben: a review of the published literature. Food Chem. Toxicol. 40(10): 1335–1373.
Stark, B. C., K. R. Pagilla and K. L. Dikshit. 2015. Recent applications of Vitreoscilla hemoglobin technology in
bioproduct synthesis and bioremediation. Appl. Microbiol. Biotechnol. 99(4): 1627–1636.
Steinbüchel, A. and S. Hein. 2001. Biochemical and molecular basis of microbial synthesis of polyhydroxyalkanoates
in microorganisms. Adv. Biochem. Eng. Biotechnol. 71: 81–123.
Tang, J., B. Liu, T. T. Chen, K. Yao, L. Zeng, C. Y. Zeng and Q. Zhang. 2018. Screening of a beta-cypermethrin
degrading bacterial strain Brevibacillus parabrevis BCP-09 and its biochemical degradation pathway.
Biodegrad. 29(6): 525–541.
Tiso, T., P. Sabelhaus, B. Behrens, A. Wittgens, F. Rosenau, H. Hayen and L. M. Blank. 2016. Creating metabolic
demand as an engineering strategy in Pseudomonas putida–Rhamnolipid synthesis as an example. Adv.
Energy Sci. Environ. Eng. 3: 234–244.
Tomasek, P. H. and J. S. Karns. 1989. Cloning of a carbofuran hydrolase gene from Achromobacter sp. strain WM111
and its expression in gram-negative bacteria. J. Bacteriol. 171(7): 4038–4044.
Tran, N. H., T. Urase, H. H. Ngo, J. Hu and S. L. Ong. 2013. Insight into metabolic and cometabolic activities of
autotrophic and heterotrophic microorganisms in the biodegradation of emerging trace organic contaminants.
Bioresour. Technol. 146: 721–731.
Turnbull, G. A., J. E. Cullington, A. Walker and J. Morgan. 2001. Identification and characterisation of a diuron
degrading bacterium. Biol. Fertil. Soils. 33: 472–476.
Unger, B. P., I. C. Gunsalus and S. G. Sligar. 1986. Nucleotide sequence of the Pseudomonas putida cytochrome
P-450cam gene and its expression in Escherichia coli. J. Biol. Chem. 261(3): 1158–1163.
Vague, M., G. Chan, C. Roberts, N. A. Swartz and J. L. Mellies. 2019. Pseudomonas isolates degrade and form
biofilms on polyethylene terephthalate (PET) plastic. bioRxiv. p: 647321.
van Hylckama Vlieg, J. E., L. Tang, J. H. LutjeSpelberg, T. Smilda, G. J. Poelarends, T. Bosma, A. E. van Merode, M.
W. Fraaije and D. B. Janssen. 2001. Halohydrin dehalogenases are structurally and mechanistically related to
short-chain dehydrogenases/reductases. J. Bacteriol. 183(17): 5058–5066.
Verma, A., K. Dhiman and P. Shirkot. 2016. Hyper-production of laccase by Pseudomonas putida LUA15. 1 through
mutagenesis. J. Microbiol. Exp. 3(1): 00080.
144 Bioremediation for Sustainable Environmental Cleanup
Volke, D. C., L. Friis, N. T. Wirth, J. Turlin and P. I. Nikel. 2020. Synthetic control of plasmid replication enables
target-and self-curing of vectors and expedites genome engineering of Pseudomonas putida. Metab. Eng.
Commun. 10: e00126.
Wallis, J. G., J. L. Watts and J. Browse. 2002. Polyunsaturated fatty acid synthesis: what will they think of next?
Trends Biochem. Sci. 27(9): 467–473.
Wang, J. P., L. X. Wu, F. Xu, J. Lv, H. J. Jin and S. F. Chen. 2010. Metabolic engineering for ethylene production
by inserting the ethylene-forming enzyme gene (efe) at the 16S rDNA sites of Pseudomonas putida KT2440.
Bioresour. Technol. 101(16): 6404–6409.
Wang, Y., C. Zhang, T. Gong, Z. Zuo, F. Zhao, X. Fan, C. Yang and C. Song. 2015. An upp-based markerless gene
replacement method for genome reduction and metabolic pathway engineering in Pseudomonas mendocina
NK-01 and Pseudomonas putida KT2440. J. Microbiol. Methods. 113: 27–33.
Watanabe, K., K. Nagahama and M. Sato. 1998. A conjugative plasmid carrying the efe gene for the ethylene-forming
enzyme isolated from Pseudomonas syringae pv. Glycinea. Phytopathol. 88(11): 1205–1209.
Whited, G. M. and D. T. Gibson. 1991. Toluene-4-monooxygenase, a three-component enzyme system that catalyzes
the oxidation of toluene to p-cresol in Pseudomonas mendocina KR1. J. Bacteriol. 173(9): 3010–3016.
WHO. 2003. Concise International Chemical Assessment Document 56:1,2,3-Trichloropropane. Geneva.
Worsey, M. J. and P. A. Williams. 1975. Metabolism of toluene and xylenes by Pseudomonas (putida (arvilla) mt-2:
evidence for a new function of the TOL plasmid. J. Bacteriol. 124(1): 7–13.
Xiao, Y., S. Chen, Y. Gao, W. Hu, M. Hu and G. Zhong. 2015. Isolation of a novel beta-cypermethrin degrading strain
Bacillus subtilis BSF01 and its biodegradation pathway. Appl. Microbiol. Biotechnol. 99(6): 2849–2859.
Yang, Y. H., C. J. Brigham, E. Song, J. M. Jeon, C. K. Rha and A. J. Sinskey. 2012. Biosynthesis of poly
(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) containing a predominant amount of 3‐hydroxyvalerate by
engineered Escherichia coli expressing propionate‐C o A transferase. 113(4): 815–823.
Yoshida, T. and N. Sethunathan. 1973. A Flat~ obacterium that degrades diazinon and parathion. Can. J. Microbiol.
19: 873–875.
Yoshida, T. and T. Nagasawa. 2007. Biological Kolbe-Schmitt carboxylation: possible use of enzymes for the direct
carboxylation of organic substrates. In: T. Matsuda [Ed.]. Future Directions in Biocatalysis. Elsevier Science.
You, I. S. and R. Bartha. 1982. Metabolism of 3, 4-dichloroaniline by Pseudomonas putida. J. Agric. Food Chem.
30(2): 274–277.
Yu, S., M. R. Plan, G. Winter and J. O. Krömer. 2016. Metabolic engineering of Pseudomonas putida KT2440 for the
production of para-Hydroxy Benzoic Acid. Front. Bioeng. Biotechnol. 4: 90.
Zhang, R., X. Xu, W. Chen and Q. Huang. 2016. Genetically engineered Pseudomonas putida X3 strain and its
potential ability to bioremediate soil microcosms contaminated with methyl parathion and cadmium. Appl.
Microbiol. Biotechnol. 100(4): 1987–1997.
Zhang, Y. X., K. Perry, V. A. Vinci, K. Powell, W. P. Stemmer and S. B. delCardayré. 2002. Genome shuffling leads
to rapid phenotypic improvement in bacteria. Nature. 415(6872): 644–646.
Zhao, J., D. Jia, J. Du, Y. Chi and K. Yao. 2019a. Substrate regulation on co-metabolic degradation of β-cypermethrin
by Bacillus licheniformis B-1. AMB Express. 9(1): 44866.
Zhao, J., X. Chen, H. X. Wei, J. Lv, C. Chen, X. Y. Liu, Q. Wen and L. M. Jia. 2019b. Nutrient uptake and utilization
in Prince Rupprecht’s larch (Larix principis-rupprechtii Mayr.) seedlings exposed to a combination of light-
emitting diode spectra and exponential fertilization. Soil Sci. Plant Nutr. 65(4): 358–368.
Zhongli, C., L. Shunpeng and F. Guoping. 2001. Isolation of methyl parathion-degrading strain M6 and cloning of the
methyl parathion hydrolase gene. Appl. Environ. Microbiol. 67(10): 4922–5.
Zylstra, G. J., W. R. McCombie, D. T. Gibson and B. A. Finette. 1988. Toluene degradation by Pseudomonas putida
F1: genetic organization of the tod operon. Appl. Environ. Microbiol. 54(6): 1498–1503.
Zylstra, G. J., S. W. Bang, L. M. Newman and L. L. Perry. 2000. Microbial degradation of mononitrophenols and
mononitrobenzoates. pp. 145–160 In: J. C. Spain, J. B. Hughes and H. J. Knackmuss [Eds.]. Biodegradation
of Nitroaromatic Compounds and Explosives. CRC Press, FL.
Chapter 9
General Aspects/Case Studies
on Sources and Bioremediation
Mechanisms of Metal(loid)s
Manoj Kumar, Sushma K. Varma, Renju,
Neeraj Kumar Singh and Rajesh Singh*
9.1 Introduction
With the growing world population, the consequential demand for industrial establishments to
suit human requirements has caused a rise in pollution in air, land and aquatic ecosystems due to
an overuse of accessible resources (Tarekegn et al. 2020). Environmental pollution by inorganic
pollutants like heavy metal(loid)s is the main problem (Kumar and Singh 2017). Apart from Arsenic
(As), Boron (B), and Selenium (Se), ‘heavy metal(loid)s’ refers to elements (metalloids and metals)
having atomic densities larger than 6 g cm–3 in general. In this category, there are both biologically
necessary (viz. Zn, Co, Mn, Cu and Cr) and non-essential components (viz. Hg, Cd and Pb). The
necessary components with nutritional value for plants and animals. As these are required only
in small quantities, are referred to as micronutrients. In the chemical industry, because they are
phytotoxins and zootoxins, non-essential metal(loid)s are referred to as “toxic elements. Higher
amounts of these non-essential metal(loids) are harmful to both plants and animals (Adriano 2001).
Heavy metal(loid)s contamination through natural and anthropogenic activities may deteriorate
the environment and cause adverse effects on humans and animals (Douay et al. 2008, Shen
et al. 2017, Liu et al. 2021, Adlane et al. 2020). Heavy metal(loid)s have become increasingly
controversial due to their possible harmful health and environmental consequences, as well as their
effects on international trade in a number of locations throughout the world. The increase of Cd in
grazing animals’ (mainly kidneys and liver) renders it unfit for human consumption; however, it
also poses a threat to the export of offal products. The Cd bioaccumulation in potatoes, wheat and
rice crops has significant consequences for the marketing of these crops on a local and international
basis (Roberts et al. 1994, Mclaughlin et al. 1996, Kirkham, 2006, Makino et al. 2006, Mavropoulos
et al. 2002, Pérez and Anderson 2009). Therefore, there is a worldwide alarm in ensuring that the
heavy metal(loid)s content of food fulfills regulatory criteria and compares well to those of other
countries. An organic amendment to soil can enhance the process of bioremediation, which is a
School of Environment and Sustainable Development, Central University of Gujarat, Gandhinagar, 382030, Gujarat, India.
* Corresponding author: [email protected]
146 Bioremediation for Sustainable Environmental Cleanup
natural phenomenon that uses fungi, bacteria and higher plants. It reduces the toxicity of soil by
changing the soil environment through bioaugmentation and biostimulation processes. The organic
amendments, higher plants and microorganisms include biological agents in the bioremediation of
contaminated soil with metal(loids). These organisms can render metal(loid)s harmless by lowering
their bioavailability and using chemical contaminants as a source of energy (Alexander 2000,
Zhuang et al. 2007).
like algae, microbes, yeast and fungi as well as in addition to ecosystems, activities and processes
facilitated by microbes (Giller et al. 2009, Wang et al. 2010, Babich and Stotzky 1985, Baath 1989).
contamination with meta(loids). The biosolid treatment has little effect on Ni and Cr contents in
soil (Illera et al. 2000). The occurrence of these metals in biosolids caused a significant rise in Zn,
Cd, Cu and Pb levels. It has been noticed that grazing and immobilization of soil are harmful to the
soil’s microflora. It can rise the concentration of metal(oids) Zn, Pb and Cu whereas decreasing C
and N in plants, animals and soil microbes (Haynes et al. 2009, Kao et al. 2006). Various metals and
metalloid sources and their toxicity to human beings are shown in Table 9.1.
9.3.1 Adsorption
Adsorption, according to Sposito (1984), is the accumulation of a solute at the interface between
a liquid solution and a solid. The process of forming a chemical connection with metal ions on
the surface of adsorbents is known as adsorption. The adsorption is categorized dominantly into
two sets, i.e., specific adsorption and non-specific (Bolan et al. 2014). By using functional groups,
specific adsorption binds the solute to the adsorbents (Sposito 1984). Solutes are bound by non
specific adsorption through electrostatic attraction (Bolan et al. 2014). The adsorbent’s properties
General Aspects/Case Studies on Sources and Bioremediation Mechanisms of Metal(loid)s 149
depend on pH, as a rise in pH attributes to a –ve charge on the adsorbent’s surface (Martı́nez-
Villegas et al. 2004). Pb precipitates to form hydroxide ions and structures hydroxyl species which is
more strongly bound in comparison to free metal ions (Martı́nez-Villegas et al. 2004). Multinuclear
metal-hydroxyl species in aqueous forms can precipitate metal hydroxyl in homogenized solutions
(Martınez-Villegas
́ et al. 2004, Park et al. 2011).
Components of the soil, such as organic matter, silicate clay, manganese oxides, iron and
aluminium, influence the metal(loid)s’ adsorption process (Merdy et al. 2009). These soil components
form inorganic and organic complexes with metal(oid)s (Bolan et al. 2003). Surface functional
groups form semi-covalent bonds with dissolved ions (Bradl 2004). The rise in pH dissociates H+ ions
from functional groups like phenolic, carboxyl, carbonyl and hydroxyl, increasing the metal cation’s
affinity (Bolan et al. 2003). The soil’s exterior surface, which has a number of hydroxyl groups,
creates stable surface complexes with heavy metalloids (Hanlie et al. 2001). Heavy metalloids form
two types of complexes with the soil: When water molecules are present between the soil’s surface
and the functional group in an outer-sphere complex, metalloids are directly attached to the functional
groups of the soil (Hanlie et al. 2001). Complexes within the inner sphere are typically more stable
than those outside (Bradl 2004). Natural ligands (fulvic, humic acid) and anthropogenic ligands
(nitrile tri-acetic acid and EDTA) form complexes with heavy metals (Bradl 2004). Phytoextraction
technology is used for the extraction of heavy metalloids by enhancing the mobility of heavy
metal(loid)-EDTA or NTA complexes (Meers et al. 2009). Numerous variables, including the type
of soil, affect the formation of metal(loid)-organic complexes, dominant cations, temperature, soil
solution pH, and ionic strength (Luo et al. 2010). pH governs negative charge formation on surface
functional groups (Yang et al. 2006). It has been noted that the soil’s constituents also contribute to
the formation of metalloid complexes (Bradl 2004). Fine particle soil has more active surface sites;
therefore, it has a high retention potential for the metal(loid) (Bradl 2004). Adsorption of metalloids
such as As(V) depends on the mineral surface characteristics like soil having low oxide content
adsorption property has a very negligible effect of increasing pH (Smith et al. 1999).
9.3.2 Methylation/Demethylation
Toxic metal(loid)s are removed by methylation through biological methods by transforming
(e.g., Hg) into methyl derivatives, which get removed further by the volatilization process
(Frankenberger and Losi 1995). Se, Hg and As methylated derivatives are acquired from biological
and chemical mechanisms. These mechanisms may reduce their toxicity by changing their volatility,
solubility and mobility. Both biological and chemical mechanisms can result in the methylation of
metal(loids) (Bolan et al. 2014). Biomethylation is a dominant process found in both aquatic and
soil environments. Biomethylation reduces toxicity by excreting methylated compounds from cells,
and are frequently volatile, e.g., organic arsenic. Researchers have categorized methylation into
two groups: fission-methylation and trans-methylation (Thayer and Brinckman 1982). The fission
of a chemical (methyl source) is known as fission methylation, not always having a methyl group
in order to completely get rid of a compound like formic acid (Bolan et al. 2014). Subsequently,
another substance that has been reduced to a methyl group captures the fission molecule. Unlike
trans-methylation, which includes the exchange of methyl groups between sources (donors) (methyl
acceptor). Microorganisms are active methylators in soil and sediments. For abiotic and biotic
methylation in sediment and soil, organic matter serves as a source of methyl-donor. The methylation
of Hg in sediments is influenced by organic matter and alternative electron acceptors (Martın- ́
Doimeadios et al. 2004). A study has shown that methylated As species result in the breakdown of
eliminating waste arsenicals or cellular organ arsenicals complex (Li et al. 2009).
Under anaerobic and aerobic conditions, methylation of Hg occurs (Martı́n-Doimeadios et al.
2004) and in undisturbed lake sediments under anaerobic conditions, significantly higher Hg
methylation was observed (Martın-Doimeadios
́ et al. 2004). Under these environments, Hg(II) ions
are methylated biologically to produce monomethyl and dimethyl compounds, which are extremely
150 Bioremediation for Sustainable Environmental Cleanup
lethal than any other form (Ullrich et al. 2001). Selenium removal from contaminated environments
by biomethylation has also been reported (Adriano et al. 2004).
9.3.4 Precipitation
In polluted soils with basic concentrations and several anions present including phosphate, carbonate,
sulfate and hydroxide, the precipitation mechanism of metal(loid)s removal has been discovered
(Ok et al. 2010). Metal(loid) precipitation like Pb and Cu with carbonate and phosphate is an
immobilization mechanism of bioremediation for elimination from soil or wastewater (McGowen
et al. 2001). Studies show that phosphate reduced the discharge of Zn, Pb and Cd (McGowen et al.
2001). Similarly, additional research presented the precipitation of Cr(III) by the addition of lime
by enhancing the soil pH (Bolan et al. 2003). The existence of iron oxyhydroxides causes changes
in the surface chemicals present on the substrate and often leads to co-precipitation of metal(loid)s
(Bolan et al. 2014, McGowen et al. 2001). The Pb(II) precipitates at pH 4 by the effect of ferric
oxyhydroxides with hydroxide chloride [Pb(OH)Cl], chloride (PbCl2) and carbonate (PbCO3),
while reacting with Mg/Al in an aqueous solution with hydroxides (Violante et al. 2007). Usually,
phosphate compounds are added to the soil to prevent heavy metal(loid) leaching (Bolan et al.
2014). Stability of metallic phosphates is found in the following order Pb > Cu > Zn (Bolan et al.
2003).
are present in microbial excreta and their derivative products (Park et al. 2011). Metallothioneins,
the cysteine-rich polypeptides bind metal(loid)s. Microbial exopolymers, metal-thiolate clusters,
cysteine-containing-glutamyl peptides and phytochelatins are associated with the reduction of
metal-(loid) binding and detoxification (Cobbett and Goldsbrough 2002, Park et al. 2011). The
bacterial species Micrococcus luteus and Azotobacter sp. immobilized Pb, i.e., 490 mg g−1 and
on a dry weight basis about 310 mg Pb g−1 on (Tornabene and Edwards 1972). Other groups of
researchers found that Sulfate-Reducing Bacteria (SRB) effectively remove Zn from the medium.
SRBs use phosphogypsum as a sulfate source for terminal electron acceptors for energy production.
Stenotrophomonas maltophilia reduced the toxicity of Se(III) from the legume rhizosphere of
Astragalus bisulcatus (Gregorio et al. 2005). The plant roots have a significant role in the chemistry
of the metalloids and soil environment (Tao et al. 2004). Metal(loid)s removal from contaminated
soils by plant roots is an emerging eco-friendly remediation technology (Ernst 1996). This
bio-transformation technology in soil depends on pH changes, microbial activity, phytochelatins,
metal binding by root exudates and plant uptake via roots. Plant root exudates form complexes with
organic acid and metal(loid)s in soil (Koo et al. 2010).
Figure 9.1. A description of the interactions between plants and metal contamination as well as potential outcomes for the
metal contaminants (Ojuederie and Babalola 2017).
metal(oid) removal also have some shortcomings. These include the requirements for nutritional
resources, certain climatic conditions, and appropriate soil properties for normal plant growth
(Karami and Shamsuddin 2010). The most significant disadvantages of phytoextraction are the
lengthy time required, which has hampered the widespread implementation of phytoremediation.
An overview of plant-metal contaminants interaction and the possible fate of the metal contaminants
is shown in Figure 9.1.
9.5.3 Phytovolatilization
Toxic metals, including mercury, selenium and arsenic, are capable of being biomethylated to
produce volatile substances that are discharged into the atmosphere. Phytovolatilization is the
mechanism which is involved in the reduction of pollutants via the transpiration of plants. The plant
absorbs the contaminant that is present in the water, passes through it, or undergoes transformation
there, then is released into the atmosphere. Water passes through the plant’s internal transport
mechanism circulating from roots up to the leaves, where the inorganics get evaporated or volatilized
and consequently released in the air enclosing the plant, potentially modifying the contaminant.
Using phytovolatilization and phytoextraction to remove metals from commercial projects is a
realistic option (Sakakibara et al. 2010). Tritium (3H), a radioactive isotope of hydrogen, has been
successfully phytovolatilized; it decomposes into a stable form of helium, having a half-life of about
12 yr. HMs can be absorbed by several plants, including Arabidopsis thaliana, Chara canescens and
Brassica juncea, converting them into their gaseous forms within plants, and then releasing them
back into the environment (Ghosh and Singh 2005). Dimethylselenides and dimethyldiselenides
are produced by plants (i.e., Brassica juncea and Arabidopsis thaliana), which are the volatile
forms of volatile Se when grown on a high Se medium. Similarly, data from a study on heavy
metal volatilization revealed that P. vittata is quite efficient at volatilizing Arsenic (As), as had been
documented by its removal by almost 90% of total intake from As-affected soils in a greenhouse
with subtropical conditions (Sakakibara et al. 2010). In contrast to the other ways of cleanup,
after toxins have been removed via volatilization, they cannot be stopped from spreading to other
areas. Similar occurrences of soil remediation based on volatilization have been recorded in many
other publications (Tangahu et al. 2011, Conesa et al. 2012). Although it is well recognized that
microbes dispense a significant function in the Se volatilization from soil systems (Karlson and
Frankenberger 1989), it was investigated that plants can fulfil the same job. B. juncea has been
recognized as a useful source for extracting Se from soils (Bauelos and Meek 1990, Baualos et al.
1993). The Se volatilization into methyl selenates has been hypothesized as a dominant mechanism
for plant Se elimination (Zayed and Terry 1994, Terry et al. 1992). Non-volatile methyl selenate
derivatives accumulate in the leaf of some plants, allowing them to extract Se from the soil. In the
Se accumulator Astragalus bisculatus, the enzyme that serves in producing methyl selenocysteine
was isolated and described (Neuhierl and Bock 1996).
General Aspects/Case Studies on Sources and Bioremediation Mechanisms of Metal(loid)s 155
9.5.4 Phytoextraction
Phytoextraction, better known as phytomining or phytoaccumulation, is the practice of raising a
certain crop that has been known for collecting toxins within its shoot system and leaves (hyper
accumulator or tolerant plant), harvesting them followed by elimination of contaminants from the
affected areas. Unlike destructive degrading mechanisms, this process produces a consolidated
plant and contaminate (mostly metallic) mass that must be disposed of or recycled. Correlated to
landfilling and excavation, this technology is based on the concentration of pollutants which leaves
a significantly smaller bulk of pollutants to be disposed off (Wani et al. 2012). Chelation is the
process by which soil-borne metal pollutants are transported by roots to tissues. By translocation
from roots to stems and leaves, metal(oids) are removed from the soil in Figure 9.2.
Figure 9.2. Metals uptake by Phytoextraction: Metal(oids) are eliminated from the soil by translocation into plant roots,
stem, and leaves (Mishra et al. 2017).
Table 9.2. Remediation of metal(loid)s by the combination of plants and microbial species.
9.6.2 Cadmium
In their analysis of 465 studies on the level of heavy metal contamination in Chinese agricultural
soils (Zhang et al. 2015); the findings revealed that heavy metals polluted around 10.18% of arable
soil, with cadmium (Cd2+) having the highest contamination rate at 7.75%. Through the food chain
and the soil-crop system, Cd2+ causes a health risk to humans (Nogawa and Kido 1996). According to
research, Cd2+ causes lesions on bone, renal problems and pulmonary inadequacy in people (Sharma
1995, Chakravarty et al. 2010). As a result, steps should be taken to reduce or neutralize its negative
effects in Cd-contaminated soil. To treat heavy metal-affected soils, microbial bioremediation is a
promising approach. Based on its 16s rRNA sequence and metabolic profile, a study was undertaken
in Wuhan, China (Xu et al. 2019). In which Raoultella sp. strain X13 was discovered to be a Cd
resistant bacterium that was isolated from a heavy metal-affected soil environment in southern
China, demonstrated Cd2+ biosorption and bioaccumulation abilities and depicted tolerance of up
to 8 mM Cd in LB medium. As a result, strain X13 has been proven to be used as a possible Cd2+
immobilizing substance in contaminated soil since it promotes plant development as well as Cd2+
fixation. In the case of Cd2+ pollution, many studies have shown that metal-resilient PGP microbes
may promote plant development, improve heavy metal stability and raise plant resistance to heavy
metals (Das et al. 2014).
9.6.3 Mercury
Mercury pollution has become a global problem, more precisely in the soil environment. It is a big
problem in mercury mining locations, artisanal gold smelting areas and chlorine industrial plants,
among other places, resulting in a lot of contamination of mercury (Driscoll et al. 2013).
The mercury mining regions of Wuchuan and Wanshan in China (Qiu et al. 2006, Li et al. 2008),
Almadén, a mercury mining zone in the Amazon, are located in the heart of the region, these regions
are mercury-prone areas (Higueras et al. 2006). In Dehua, Fujian Province, China, small-scale and
artisanal gold mining activities occur, and an abandoned gold mining and smelting area (United
Nations Environment Programme 2013) are all examples of mercury-polluted locations that cause
serious ecological and environmental risks. Remediating mercury-polluted sites is an imminent
but pressing task. In an investigation carried out in Beijing, China (Xun et al. 2017), Cyrtomium
macrophyllum was naturally growing in the soil mercury region. Cyrtomium macrophyllum
demonstrated remarkable mercury accumulation and translocation capabilities (Xun et al. 2017).
techniques have indefinite use, like the production of biodiesel, bioethanol, CO2 fixation, heavy
metals pollution controls and so on. And it is also favorable to treat wastewater and polluted land
because the contaminated site is highly nutrient-rich for some organisms like Ideonella sakaiensis,
a plastic waste degrading bacterium, Sulfurospirillum arsenophilum, Bacillus arsenicoselenatis,
Chrysiogenes arsenatis and Archaea (Pyrobaculum arsenaticum, Pyrobaculum aerophilum) used
for arsenic pollution control of polluted water (/Kudo et al. 2013). Species like Brassica napus can
even control soil heavy metals pollution by cadmium, cobalt and nickel (Boros-Lajszner et al. 2021).
Bioremediation is not a new technique to treat pollution load, but it is a slow process. But from
the future prospective, highly efficient bioremediation techniques need to be developed by advanced
research with the help of bioinformatics, biostatics analyses and other omics approaches helping to
understand the metabolism of microorganisms efficiently. Certain significant aspects that should be
taken care of include: (1) net reduction of pollution in actual and laboratory conditions, that requires
better equipment handling, patience and regular monitoring; (2) studying the possible contributing
factors (abiotic and biotic); and (3) the remediation technology should be cost-effective, reliable
and rapid. Moreover, the endeavor should be made to integrate the phytoremediation method with
bio-energy for the two-fold utilization of plants for phytoremediation and bio-fuel generation on
polluted lands. These methods should be beneficial to the phytoremediation of polluted regions
and concurrently generate sustainable power that can balance the prices that bear on this kind of
methodologies (Mosa et al. 2016).
9.8 Conclusion
Being an eco-friendly technology for heavy metals remediation and organic contaminants,
phytoremediation has a lot of promise. Rhizospheric bacteria and plants have shown the capability
of detoxifying and converting organic pollutants into harmless compounds that may be eliminated
from the soil without accumulating. The above-ground biomass can carry toxic metals, and
plants can then detoxify, translocate, accumulate and recover heavy metals like lead. Although
phytoremediation has excellent potential to be applied to contaminant removal from water,
sediment and soil, it has not been widely commercialized or implemented on a broad basis. Field
implementation of phytoremediation has still not been studied A lack of comprehensive insight into
the uptake mechanism of metals from the soil system to the roots has hampered the commercialization
of phytoremediation for heavy metals and metalloids. Several recent studies have focused on
transcriptomic and proteomic techniques for metal accumulation in plants and elucidating heavy
metal transport. This chapter summarizes the efficient, environmentally sound and cost-efficient
methods for removing metal(oids) from the environment and protecting the ecology.
Acknowledgment
All the authors listed in the chapter have contributed substantially, directly and intellectually to the
work and have approved its publication. The UGC, GOI, New Delhi, is sincerely appreciative of its
support of the NFSC fellowship on behalf of one of the authors, Kumar M.
References
Abhilash, P. C., J. R. Powell, H. B. Singh and B. K. Singh. 2012. Plant-microbe interactions: novel applications for
exploitation in multipurpose remediation technologies. Trends Biotechnol. 30: 416–420.
Adediran, G. A., B. T. Ngwenya, J. F. W. Mosselmans, K. V. Heal and B. A. Harvie. 2015. Mechanisms behind
bacteria induced plant growth promotion and Zn accumulation in Brassica juncea. J. Hazard. Mater. 283:
490–499.
Adlane, B., Z. Xu, X. Xu, L. Liang, J. Han and G. Qiu. 2020. Evaluation of the potential risks of heavy metal
contamination in rice paddy soils around an abandoned Hg mine area in Southwest China. Acta Geochim.
39: 85–95.
160 Bioremediation for Sustainable Environmental Cleanup
Adriano, D. C., J. Albright, F. W. Whick, I. K. Islandar and C. Sherony. 1997. Remediation of soil contaminated with
metals and radionuclide-contaminated soils. pp. 27–46. In: I. K. Islandar and D. C. Adriano (Eds.). Remediation
of Soils Contaminated with Metals. Northwood, UK. SciRev-Review the scientific review process.
Adriano, D. C. 2001. Bioavailability of trace metals. In: Trace Elements in Terrestrial Environments Springer, New
York.
Adriano, D. C., W. W. Wenzel, J. Vangronsveld and N. S. Bolan. 2004. Role of assisted natural remediation in
environmental cleanup. Geoderma. 2-4: 121–142.
Agnello, A. C., M. Bagard, E. D. van Hullebusch, G. Esposito and D. Huguenot. 2016. Comparative bioremediation
of heavy metals and petroleum hydrocarbons co-contaminated soil by natural attenuation, phytoremediation,
bioaugmentation and bioaugmentation-assisted phytoremediation. Sci. Total Environ. 563: 693–703.
Alexander, M. 2000. Aging, bioavailability, and overestimation of risk from environmental pollutants. Environ. Sci.
Technol. 34: 4259–4265.
Alloway, B. J. 2013. Sources of heavy metals and metalloids in soils. In: B. Alloway (Eds.). Heavy Metals in Soils:
Environmental Pollution Springer, Dordrecht.
Ando, J. 1987. Thermal phosphate. pp. 93–124. In: F. T. Nielsson (Eds.). Manual of Fertilizer Processing, Fertilizer
Science and Technology Series. New York, Marcel Dekker Inc.
Anjaneya, O., S. Y. Souche, M. Santoshkumar and T. B. Karegoudar. 2011. Decolorization of sulfonated azo dye
Metanil Yellow by newly isolated bacterial strains: Bacillus sp. strain AK1 and Lysinibacillus sp. strain AK2.
J. Hazard. Mater. 190(1): 351–358.
Anju, M. and D. K. Banerjee. 2003. Heavy metal levels and solid phase speciation in street dusts of Delhi, India.
Environ. Pollut. 123: 95–105.
Anju, M. and D. K. Banerjee. 2010. Comparison of two sequential extraction procedures for heavy metal partitioning
in mine tailings. Chemosphere. 78: 1393–1402.
Anju, M. and D. K. Banerjee. 2011. Associations of cadmium, zinc, and lead in soils from a lead and zinc mining area
as studied by single and sequential extractions. Environ. Monit. Assess. 176: 67–85.
Baath, E. 1989. Effects of heavy metals in soil on microbial processes and populations (a review). Water Air Soil
Pollut. 47: 335–379.
Babich, H. and G. Stotzky. 1985. Heavy metal toxicity to microbe-mediated ecologic processes: a review and potential
application to regulatory policies. Environ. Res. 36: 111–37.
Babu, A. G., P. J. Shea, D. Sudhakar, I. BooJung and B. T. Oh. 2015. Potential use of Pseudomonas koreensis
AGB-1 in association with Miscanthus sinensis to remediate heavy metal(loid)-contaminated mining site soil.
J. Environ. Manage. 151: 160–166.
Bachate, S. P., Khapare, R. M. Kisan and M. K. Kodam. 2012. Oxidation of arsenite by two β-proteobacteria isolated
from soil. Appl. Microbiol. Biotechnol. 93: 2135–2145.
Baker, A. and R. Brooks. 1989. Terrestrial higher plants which hyper accumulate metallic elements—a review of their
distribution, ecology and phytochemistry. Biorecovery. 1: 81–126.
Battaglia‐Brunet, F., M. C. Dictor, F. Garrido, C. Crouzet, D. Morin and K. Dekeyser. 2002. An arsenic (III)‐oxidizing
bacterial population: selection, characterization, and performance in reactors. J. Appl. Microbiol. 93: 656–6.
Bauelos, G. and D. Meek. 1990. Accumulation of selenium in plants grown on selenium-treated soil. J. Environ. Qual.
19: 772.
Bernd, L. 2007. Characterization, Treatment and Environmental Impacts. In Mine Wastes.: Springer, Berlin,
Heidelberg.
Baualos, G. S., G. Cardon, B. Mackey, J. Ben-Asher, L. Wu, P. Beuselinck, S. Akohoue and S. Zambrzuski. 1993.
Plant and environment interactions, boron and selenium removal in boron-laden soils by four sprinkler
irrigated plant species. J. Environ. Quali. 22: 786–792.
Bittsanszkya. A., T. Kömives, G. Gullner, G. Gyulai, J. Kiss, L. Heszky, L. Radimszky and H. Rennenberg. 2005.
Ability of transgenic poplars with elevated glutathione content to tolerate zinc (2+) stress. Environ. Int.
31: 251–254.
Blaylock, M. J., D. E. Salt, S. Dushenkov, O. Zakharova, C. Gussman, Y. Kapulnik, B. D. Ensley and I Raskin.
1997. Enhanced accumulation of Pb in Indian mustard by soil-applied chelating agents. Environ. Sci. Technol.
31: 860–865.
Bolan, N., A. Kunhikrishnan, R. Thangarajan, J. Kumpiene, J. Park and T. Makino. 2014. Remediation of heavy
metal(loid)s contaminated soils-To mobilize or to immobilize. J. Hazard. Mater. 266: 141–166.
Bolan, S., C. Adriano and R. Naidu. 2003. Role of phosphorus mobilization and bioavailability of heavy metals in the
soil-plant system. Rev. Environ. Contam. Toxicol. 177: 1–44.
Boros-Lajszner, E., J. Wyszkowska, A. Borowik and J. Kucharski 2021. The response of the soil microbiome to
contamination with cadmium, cobalt and nickel in soil sown with Brassica napus. Minerals. 11: 498.
Bradl, H. 2004. Adsorption of heavy metal ions on soils and soils constituents. J. Colloid Interface Sci. 277: 1–18.
General Aspects/Case Studies on Sources and Bioremediation Mechanisms of Metal(loid)s 161
Geller, W., H. Klapper and M. Schultze. 1998. Natural and anthropogenic sulfuric acidification of lakes. In: W. Geller,
H. Klapper and W. Salomons [eds.]. Acidic Mining Lakes: Acid Mine Drainage, Limnology and Reclamation.
Berlin, Heidelberg: Springer.
Ghosh, M. and S. P. Singh. 2005. A review on phytoremediation of heavy metals and utilization of it’s by products.
Appl. Ecol. Environ. Sci. 3: 1–18.
Giller, E., E. Witter and S. P. McGrath. 2009. Heavy metals and soil microbes. Soil Biol. Biochem. 41: 2031–2037.
Gove, L., C. M. Cooke, F. A. Nicholson and A. J. Beck. 2001. Movement of water and heavy metals (Zn, Cu, Pb
and Ni) through sand and sandy loam amended with biosolids under steady-state hydrological conditions.
Bioresour. Technol. 78: 171–9.
Gregorio, S. D., S. Lampis and G. Vallini. 2005. Selenite precipitation by a rhizospheric strain of Stenotrophomonas
sp. isolated from the root system of Astragalus bisulcatus: a biotechnological perspective. Environ. Int. 31:
233–241.
Gupta, P. and V. Kumar. 2017. Value added phytoremediation of metal stressed soils using phosphate solubilizing
microbial consortium. World J. Microbiol. Biotechnol. 9: 33.
Hanlie, H., F. Zhengyi and M. Xinmin. 2001. The adsorption of [Au(Hs)2]—On Kaolinite Surfaces: Quantum
Chemistry Calculations. Canad Mineral. 39: 1591–1596.
Haruma, T., K. Yamaji, K. Ogawa, H. Masuya, Y. Sekine and N. Kozai. 2019. Root-endophytic Chaetomium cupreum
chemically enhances aluminium tolerance in Miscanthus sinensis via increasing the aluminium detoxicants,
chlorogenic acid and oosporein. PLoS One. 14: 0212644.
Haynes, R. J., G. Murtaza and R. Naidu. 2009. Inorganic and organic constituents and contaminants of biosolids:
implications for land application. Adv. Agron. 104: 165–267.
Higueras, P., R. Oyarzun, J. Lillo, J. Sánchez-Hernandez, J. Molina, J. Esbri and S. Lorenzo. 2006. The Almadén
district (Spain): anatomy of one of the world’s largest Hg-contaminated sites. Sci. Total Environ. 356: 112–124.
Holden, J. F. and M. Adams. 2003. Microbe–metal interactions in marine hydrothermal environments. Curr. Opin.
Chem. Biol. 7: 160–165.
Hsu, N. H., S. L. Wang, Y. C. Lin, G. D. Sheng and J. F. Lee. 2009. Reduction of Cr(VI) by crop-residue derived black
carbon. Environ. Sci. Technol. 43: 8801–8806.
Huang, J. W., J. Chen and W. R. Berti. 1997. Phytoremediation of lead-contaminated soils: role of synthetic chelates
in lead phytoextraction. Environ. Sci. Technol. 31: 800–805.
Iimura,Y., M. Yoshizumi, T. Sonoki, M. Uesugi, K. Tatsumi, K. Horiuchi, S. Kajita and Y. Katayama. 2007. Hybrid
aspen with a transgene for fungal manganese peroxidase is a potential contributor to phytoremediation of the
environment contaminated with bisphenol. J. Wood Sci. 53: 541–544.
İlay, R., A. Baba and Y. Kavdır. 2019. Removal of metals and metalloids from acidic mining lake (AML) using olive
oil solid waste (OSW). Int. J. Environ. Sci. Technol. 16: 4047–4058.
Illera, V., I. Walter, P. Souza and V. Cala. 2000. Short-term effects of biosolid and municipal solid waste applications
on heavy metals distribution in a degraded soil under a semi-arid environment. Sci. Total Environ. 255: 29–44.
Jadia, C. D. and M. H. Fulekar. 2009. Phytoremediation of heavy metals: recent techniques. Afr. J. Biotechnol. 8:
921–928.
Jin, T., C. Shi, P. Wang, J. Liu and L. Zhan. 2021. A review of bioremediation techniques for heavy metals pollution
in soil. In IOP Conference Series: Environ. Earth Sci. IOP Publishing. 012012.
Jin, Z., S. Deng, Y. Wen, Y. Jin, L. Pan and Y. Zhang, T. Black, K. C. Jones, H. Zhang and D. Zhang. 2019. Application
of Simplicillium chinense for Cd and Pb biosorption and enhancing heavy metal phytoremediation of soils. Sci.
Total Environ. Sci. 697: 134148.
Jinadasa, K. B. P. N., P. J. Milha, C. A. Hawkins, P. S. Cornish, P. A. Williams and C. J. Kaldo. 1997. Heavy Metals
in the Environment. Richmond, Australia: University of Western Sydney, Hawkesbury Campus, Faculty of
Agriculture and Horticulture.
Kao, P. H. C. C. Huang and Z. Y. Hseu. 2006. Response of microbial activities to heavy metals in a neutral loamy soil
treated with biosolid. Chemosphere. 64: 63–70.
Kapahi, M. and S. Sachdeva. 2019. Bioremediation options for heavy metal pollution. J. Health Pollut. 24: 191203.
Karami, A. and Z. H. Shamsuddin. 2010. Phytoremediation of Heavy Metals with Several Efficiency Enhancer
Methods. 9: 3689–3698.
Kärenlampi, S., H. Schat, J. Vangronsveld, J. A. C. Verkleji, D. van Der Lelei, M. Mergeay and A. I. Tervahauta
2000. Genetic engineering in the improvement of plants for phytoremediation of metal polluted soils. Environ.
Pollut. 107: 22.
Karlson, U. and W. J. Frankenberger. 1989. Accelerated rates of selenium volatilization from California soils. Soil
Sci. Soc. Am. J. 53: 749–753.
Kim, J. G., J. B. Dixon and C. C. Chusuei. 2002. Oxidation of chromium(III) to (VI) by manganese oxides. Soil Sci
Soc Am J. 66: 306–315.
General Aspects/Case Studies on Sources and Bioremediation Mechanisms of Metal(loid)s 163
Kirkham, M. B. 2006. Cadmium in plants on polluted soils: effects of soil factors, hyperaccumulation, and amendments
soil factors, hyperaccumulation, and amendments. Geoderma. 137: 19–32.
Koo, B. J., W. Chen, A. C. Chang, A. L. Page, T. C. Granato and R. H. Dowdy. 2010. A root exudates based approach to
assess the long-term phytoavailability of metals in biosolids-amended soils. Environ. Pollut. 158: 2582–2588.
Kotrba, P., L. Dolecková, V. de Lorenzo and T. Ruml. 1999. Enhanced bioaccumulation of heavy metal ions by
bacterial cells due to surface display of short metal binding peptides. Appl. Environ. Microbiol. 65: 1092–1098.
Krämer, U., J. D. Cotter-Howells, J. M. Charnock, A. J. M. Baker and J. A. C. Smith. 1996. Free histidine as a metal
chelator in plants that accumulate nickel. Nature. 379: 635–638.
Krämer, U. 2005. Phytoremediation: novel approaches to cleaning up polluted soils. Curr. Opin. Biotechnol. 16:
133–141.
Krebs, W., C. Brombacher, P. P. R. Bosshard Bachofen and H. Brandl. 1997. Microbial recovery of metals from solids.
FEMS Microbiol. Rev. 20: 605–617.
Kudo, K., N. Yamaguchi, T. Makino, T. Ohtsuka, K. Kimura, D. T. Dong and S. Amachi. 2013. Release of arsenic
from soil by a novel dissimilatory arsenate reducing bacterium, Anaeromyxobacter sp. Strain PSR-1. Appl.
Environ. Microbiol. 79: 4635–4642.
Kumar, M. and R. Singh. 2017. Phytoremediation: a green technology for remediation of metal-contaminated sites.
pp. 306–328. In: R. N. Bharagava (Ed.). Environmental Pollutants and their Bioremediation Approaches. CRC
Press, Boca Raton, Florida.
Kumar, P. and M. H. Fulekar. 2018. Research article rhizosphere bioremediation of heavy metals (copper and lead) by
Cenchrus ciliaris. J. Environ. Sci. 12: 166–176.
Kumar, S., J. Dubey, S. Mehra, P. Tiwari and A. J. Bishwas. 2011. Potential use of cyanobacterial species in
bioremediation of industrial effluents. Afr. J. Biotechnol. 10: 1125–1132.
Lambert, M., G. Pierzynski, L. Erickson and J. Schnoor. 1994. Remediation of lead, zinc, and cadmium contaminate.
Environ. Sci. Technol. 7: 91–102.
Lasat, M. M., A. J. Baker and L. V. Kochian. 1998. Kochian Altered Zn compartmentation in the root symplasm
and stimulated Zn absorption into the leaf as mechanisms involved in Zn hyperaccumulation in Thlaspi
caerulescens. Plant Physiol. 118: 875–883.
Letterman, R. D. and W. J. Mitsch. 1978. Impact of mine drainage on a mountain stream in Pennsylvania. Environ.
Pollut. 17: 53–73.
Li, P., X. Feng, G. Qiu, L. Shang and S. Wang. 2008. Mercury exposure in the population from 568 Wuchuan mercury
mining area, Guizhou, China. Sci. Total Environ. 395: 72–79.
Li, R. Y., Y. Ago, W. J. Liu, N. Mitani, J. Feldmann, S. P. McGrath, J. F. Ma and F. J. Zhao. 2009. The rice Aquaporin
Lsi1 mediates uptake of methylated Arsenic species. Plant Physiology. 150: 2071–2080.
Li, X., W. Peng, Y. Jia, L. Lu and W. Fan. 2016. Bioremediation of lead contaminated soil with Rhodobacter
sphaeroides. Chemosphere. 156: 228–235.
Lindström, E. B. and H. M. Sehlin. 1989. High efficiency of plating of the thermophilic sulfur-dependent
archaebacterium Sulfolobus acidocaldarius. Appl. and Environ. Microbiol. 55: 3020–3021.
Liu, Y., C. J. Lin, W. Yuan, Z. Lu and X. Feng. 2021. Translocation and distribution of mercury in biomasses from
subtropical forest ecosystems: evidence from stable mercury isotopes. Acta Geochim. 40: 42–50.
Lone, M. I., Z. He, P. J. Stoffella and X. Yang. 2008. Phytoremediation of heavy metal polluted soils and water:
progresses and perspectives. J. Zhejiang Univ. Sci. 9: 210–220.
Luo, Z., A. Wadhawan and E. J. Bouwer. 2010. Sorption behavior of nine chromium (III) organic complexes in soil.
Int. J. Environ. Sci. Technol. 7: 1–10.
Ma, Y., C. Zhang, R. S. Oliveira, H. Freitas and Y. Luo. 2016. Bioaugmentation with endophytic bacterium E6S
homologous to Achromobacter piechaudii enhances metal rhizoaccumulation in host Sedum plumbizincicola.
Front. Plant Sci. 7: 1–9.
Ma, Y., M. Rajkumar, R. S. Oliveira, C. Zhang and H. Freitas. 2019. Potential of plant beneficial bacteria and
arbuscular mycorrhizal fungi in phytoremediation of metal-contaminated saline soils. J. Hazard. Mater. 379:
120813.
Maestri, E. and M. Marmiroli. 2011. Genetic and molecular aspects of metal tolerance and hyperaccumualtion. Metal
Toxicity in Plants: Perception, Signalling and Remediation, 41–61.
Mahmood, R., F. Sharif, S. Ali and M. U. Hayyat. 2013. Bioremediation of textile effluent by indigenous bacterial
consortia and its effects on Zea mays L. CV C1415. J. Anim. Plant Sci. 23: 1193–1199.
Makino, T. K., K. Sugahara, Y. Sakurai, H. Takano, T. Kamiya and K. Sasaki. 2006. Remediation of cadmium
contamination in paddy soils by washing with chemicals: selection of washing chemicals. Environ. Pollut.
144: 2–10.
164 Bioremediation for Sustainable Environmental Cleanup
Martın-Doimeadios,
́ R. R., E. Tessier, D. Amouroux, R. Guyoneaud, R. Duran, P. Caumette and O. F. X. Donard.
2004. Mercury methylation/demethylation and volatilization pathways in estuarine sediment slurries using
species-specific enriched stable isotopes. Mar. Chem. 90: 107–123.
Martınez-Villegas,
́ N., L. M. Flores-Vélez and O. Domınguez. 2004. Sorption of lead in soil as a function of pH: a
study case in México. Chemosphere. 57: 1537–1542.
Mateos, L. M., E. Ordóñez, M. Letek and J. A. Gil. 2006. Corynebacterium glutamicum as a model bacterium for the
bioremediation of arsenic. Int. J. Microbiol. 3: 207–215.
Mavropoulos, E., A. M., Rossi, A. M. Costa, C. A. C. Perez, J. C. Moreira and M. Saldanha. 2002. Studies on the
mechanisms of lead immobilization by hydroxyapatite. Environ. Sci. Technol. 36: 1625–1629.
McBride, M. B. 1995. Toxic metal accumulation from agricultural use of sludge Are USEPA regulations protective.
J. Environ. Qual. 24: 5–18.
McGowen, S. L., N. T. Basta and G. O. Brown. 2001. Use of diammonium phosphate to reduce heavy metal solubility
and transport in smelter-contaminated soil. J. Environ. Qual. 30: 493–500.
Mclaughlin, M. J., K. G. Tiller, R. Naidu and D. P. Stevens. 1996. Review: the behaviour and environmental impact
of contaminants in fertilizers. Aust. J. Soil Res. 34: 1–54.
Meers, E., M. Qadir, P. De Caritat, F. M. G. Tack, G. Du Laing and M. H. Zia. 2009. EDTA-assisted Pb
phytoextraction. Chemosphere. 74(10): 1279–1291.
Merdy, P., T. Gharbi and Y. Lucas. 2009. Pb, Cu and Cr interactions with soil: Sorption experiments and modelling.
Colloids Surf. A: Physicochem. Eng. Asp. 347: 192–199.
Milner. M. J. and V. Kochian. 2008. Investigating heavy-metal hyperaccumulation using Thlaspi caerulescens as a
model system. Ann. Bot. 102: 3–13.
Mishra, J., R. Singh and N. K. Arora. 2017 Alleviation of heavy metal stress in plants and remediation of soil by
rhizosphere microorganisms. Front. Microbiol. 8: 1706.
Monica, M. O., N. W. Julia and M. M. Raina. 2008. Characterization of a bacterial community in an abandoned
semiarid lead-zinc mine tailing site. Appl. Environ. Microbiol. 74: 3899–3907.
Mosa, K. A., I. Saadoun, K. Kumar, M. Helmy and O. P. Dhankhe. 2016. Potential biotechnological strategies for the
cleanup of heavy metals and metalloids. Front. Plant Sci. 7: 303.
Mukhopadhyay, S. and S. K. Maiti. 2010. Phytoremediation of metal mine waste. Appl. Ecol. Environ. Sci.
8: 207–222.
Nagata, S., K. Yamaji, N. Nomura and H. Ishimoto. 2015. Root endophytes enhance stress-tolerance of Cicuta virosa L.
growing in a mining pond of eastern Japan. Plant Species Biol. 30: 116–125.
Nandakumar, P., V. Dushenkov, H. Motto and I. Raskin. 1995. Phytoextraction: The use of plants to remove heavy
metals from soils. Environ. Sci. Technol. 29: 1232–1238.
Nayak, A. K., S. S. Panda, A. Basu and N. K. Dhal. 2018. Enhancement of toxic Cr (VI), Fe, and other heavy metals
phytoremediation by the synergistic combination of native Bacillus cereus strain and Vetiveria zizanioides L.
Int. J. Phytoremediation. 20: 682–691.
Nejad, Z. D., M. C. Jung and K. H. Kim. 2018. Remediation of soils contaminated with heavy metals with an
emphasis on immobilization technology. Environ. Geochem. Health. 40: 927–953.
Neuhierl, B. and A. Bock. 1996. On the mechanism of selenium tolerance in selenium-accumulating plants:
purification and characterization of a specific selenocysteine methyltransferase from cultured cells of
Astragalus bisculatus. Eur. J. Biochem. 239: 235–238.
Nicholson, F. A., B. J. Chambers, J. R. Williams and R. J. Unwin. 1999. Heavy metal contents of livestock feeds and
animal manures in England and Wales. Bioresour. Technol. 70: 23–31.
Nogawa, K. and T. Kido. 1996. Itai-Itai disease and health effects of cadmium. Toxicology of Metals, 353–369.
Ojuederie, O. B. and O. O. Babalola. 2017. Microbial and plant-assisted bioremediation of heavy metal polluted
environments: a review. Int. J. Environ. Res. Public Health. 14: 1504.
Ok, Y. S., S. E. Oh, M. Ahmad, S. Hyun, K. R. Kim and D. H. Moon. 2010. Effects of natural and calcined oyster
shells on Cd and Pb immobilization in contaminated soils. Environ. Earth Sci. 61: 1301–8.
Ordóñez, E., M. Letek, N. Valbuena, J. A. Gil and L. M. Mateos. 2005. Analysis of genes involved in arsenic resistance
in Corynebacterium glutamicum ATCC 13032. Appl. Environ. Microbiol. 71: 6206–6215.
Park, J. H., D. Lamb, P. Paneerselvam, G. Choppala, N. Bolan and J. W. Chung. 2011. Role of organic amendments
on enhanced bioremediation of heavy metal(loid). J. Hazard. Mater. 185: 549–574.
Pérez, A. L. and K. A. Anderson. 2009. DGT estimates cadmium accumulation in wheat and potato from phosphate
fertilizer applications. Sci. Total Environ. 407: 5096–103.
Pierzynski, G. M. 1997. Strategies for remediating trace-element contaminated sites. pp. 67–84. In: I. K. Iskandar and
D. C. Adriano [Eds.]. Remediation of Soils Contaminated with Metals. Advances in Environmental Science.
Middlesex, UK. Science Reviews.
Pointing, S. B. 2001. Feasibility of bioremediation by white-rot fungi. Appl. Microbiol. Biotechnol. 57: 20–33.
General Aspects/Case Studies on Sources and Bioremediation Mechanisms of Metal(loid)s 165
Umrania, V. 2006. Bioremediation of toxic heavy metals using acidothermophilic autotrophes. Bioresour. Technol.
97: 1237–1242.
United Nations Environment Programme. 2013. Global Mercury Assessment Sources, Emissions, Releases, and
Environmental Transport. Geneva, Switzerland: United Nations Environment Programme Chemicals Branch.
Valls, M., S. Atrian, V. de Lorenzo and L. A. Fernández. 2000. Engineering a mouse metallothionein on the cell
surface of Ralstonia eutropha CH34 for immobilization of heavy metals in soil. Nat. Biotechnol. 18: 661–665.
Aken, B. V., R. Tehrani and J. L. Schnoor. 2011. Endophyte-assisted phytoremediation of explosives in poplar trees
by Methylobacterium populi BJ001 T. In Endophytes of Forest Trees (pp. 217–234). Springer, Dordrecht.
Van Dillewijn, P., J. L. Couselo, E. Corredoira, A. Delgado, A. Ballester and J. L. Ramo. 2008 Bioremediation of
2, 4, 6-trinitrotoluene by bacterial nitroreductase expressing transgenic aspen. Environ. Sci. Technol. 42:
7405–7410.
Violante, A., S. D. Gaudio, M. Pigna, M. Ricciardella and D. Banerjee. 2007. Coprecipitation of arsenate with metal
oxides. 2. Nature, Mineralogy, and Reactivity of Iron(III) Precipitates. Environ. Sci. Technol. 41: 8275–8280.
Wang, F., J. Yao, Y. Si, H. Chen, M. Russel and K. Chen. 2010. Short-time effect of heavy metals upon microbial
community activity community activity. J. Hazard. Mater. 15: 510–6.
Wani, S. H., G. S. Sanghera, H. Athokpam, J. Nongmaithem, R. Nongthongbam, B. S. Naorem and H. S. Athokpam.
2012. Phytoremediation: curing soil problems with crops. Afr. J. Agric. Res. 7: 3991–4002.
Wiatrowski, H. A., P. M. Ward and T. Barkay. 2006. Novel reduction of mercury (II) by mercury-sensitive dissimilatory
metal-reducing bacteria. Environ. Sci. Technol. 40: 6690–6696.
Wixson, G. and B. Davies. 1994. Guidelines for lead in soil: proposal of the society of environmental geochemistry
and health. Environ. Sci. Technol. 28: 26A–31A.
Xiong, X., L. Yanxia, L. Wei, L. Chunye H. Wei and Y. Ming. 2010. Copper content in animal manures and potential
risk of soil copper pollution with animal manure use in agriculture. Resour. Conserv. Recycl. 54: 985–990.
Xu, S., Y. Xing, S. Liu, Q. Huang and W. Chen. 2019. Chen Role of novel bacterial Raoultella sp. strain X13 in plant
growth promotion and cadmium bioremediation in soil. Appl. Microbiol. Biotechnol. 103: 3887–3897.
Xun, Y., L. Feng, Y. Li and H. Dong. 2017. Mercury accumulation plant Cyrtomium macrophyllum and its potential
for phytoremediation of mercury polluted sites. Chemosphere. 189: 161–170.
Yamaji, K., Y. Watanabe, H. Masuya, A. Shigeto, H. Yui and T. Haruma. 2016. Root fungal endophytes enhance
heavy-metal stress tolerance of Clethra barbinervis growing naturally at mining sites via growth enhancement,
promotion of nutrient uptake and decrease of heavy-metal concentration. PLoS One. 11: 1–15.
Yang, J. Y., X. E. Yang, Z. L. He, T. Q. Li, J. L. Shentu and P. J. Stoffella. 2006. Effects of pH, organic acids, and
inorganic ions on lead desorption from soils. Environ. Pollut. 143: 9–15.
Ok, Y. S., S. E. Oh, M. Ahmad, S. Hyun, K. R. Kim and D. H. Moon. 2010. Effects of natural and calcined oyster
shells on Cd and Pb immobilization in contaminated soils. Environ. Earth Sci. 61: 1301–8.
Zayed, A. M. and N. Terry. 1994. Selenium volatilization in roots and shoots: effects of shoot removal and sulfate
level. J. Plant Physiol. 143: 8–14.
Zhang, M. Y., A. Bourbouloux, O. Cagnac, C. V. Srikanth, D. Rentsch, A. K. Bachhawat and S. Delrot. 2004. A novel
family of transporters mediating the transport of glutathione derivatives in plants. Plant Physiol. 134: 482–491.
Zhang, X., T. Zhong and L. Liu. 2015. Ouyang X. Impact of soil heavy metal pollution on food safety in China. PLoS
One. 10: 1–14.
Zhuang, X., J. Chen, H. Shim and Z. Bai. 2007. New advances in plant growth-promoting rhizobacteria for
bioremediation. Environ. Int. 33: 406–13.
Złoch, M., C. T. Kowalkowski, J. Tyburski and K. Hrynkiewicz. 2017. Modeling of phytoextraction efficiency
of microbially stimulated Salix dasyclados L. in the soils with different speciation of heavy metals Int. J.
Phytoremediation. 19: 1150–1164.
Chapter 10
Metal(loid)s Toxicity
and Bacteria Mediated
Bioremediation
Sushant Sunder,1 Anshul Gupta,1 Mehak Singla,1
Rohit Ruhal 2 and Rashmi Kataria1,*
10.1 Introduction
Pollutants are impurities that enter the natural environment beyond permissible limits and have
a notable negative impact on the inhabitants. These chemical compounds could be new to the
environment or present naturally, such as metal pollution in mining regions, volcanic eruptions, air
fallout, sea salt particles, etc. Industrialization and aggressive utilization of natural resources have
produced a number of contaminants or pollutants that are either biodegradable or non-biodegradable
in nature. Contaminants including toxic metals, heavy metals, insecticides, herbicides and
radioactive compounds are among the pollutants that have caused significant environmental and
health problems. It has been observed that pollutants are being continuously released from industrial
waste and anthropogenic sources to the environment, and this waste circulates from the physical
environment to plants, plants to animals, and so on enter the food chain to reach humans.
1
Department of Biotechnology, Delhi Technological University (DTU), Bawana Road, Delhi 110042.
2
Department of Biological Sciences. 1250 W Wisconsin Avenue, Marquette University, Milwaukee, Wisconsin 53233,
USA.
* Corresponding author: [email protected]
168 Bioremediation for Sustainable Environmental Cleanup
metals such as beryllium may be hazardous. However, the necessary elements or metals, such as
iron (Fe), could be poisonous. Metals in certain oxidation phases may also be toxic, such as Cr (III) is
an essential micronutrient, whilst the other form Cr (VI) is a potential carcinogen. The consequences
of heavy metal toxicity include decreased fitness, problems with reproduction, cancers and mortality.
Organometallic forms, such as methyl Hg (mercury) and tetraethyl Pb, may be extremely hazardous,
while organometallic derivatives, such as cobaltocenium cation, are less toxic. Radium, which is a
radioactive heavy metal, mimics the property of calcium as it gets absorbed in bones and, along with
Pb and Hg, causes several health complications. However, there are exceptions in cases of Barium
(Ba) and Aluminum (Al), which are rapidly eliminated by the kidneys.
concluded that lead (Pb) could influence behavioural inhibition systems, resulting in an increase of
aggression, and it could also promote tooth decay.
The rate of consumption is also a determining step for toxicity in human health. Metals such
as copper, nickel and zinc are essential trace elements for the body but are harmful if consumed
in excess. Prolonged exposure to copper is linked to abnormalities that severely affect capillaries,
kidneys and liver functions. It also leads to central nervous system irritation along with depression.
System dysfunction impedes growth and development. Excess zinc is related to infertility. Exposure
to nickel among humans enhances the risk of developing respiratory cancer nasopharyngeal
carcinoma (Barta et al. 2019). Excessive mercury harms the neurological system and affects
muscle functioning, body balance, partial blindness and abnormalities in infants (Baby et al. 2010).
According to some studies, mercury may harm both the foetal and embryonic neural systems at
concentrations lower than the WHO’s recommended limits, resulting in learning difficulties, poor
memory and reduced attention spans.
Cadmium (Cd) poisoning damages the liver, brain, lungs, placenta, kidneys and bones. The
severity depends on the amount of exposure (Sobha et al. 2007). Extreme exposure has been
related to mortality and pulmonary oedema. Emphysema, bronchiolitis and alveolitis are the other
consequences of exposure to Cd. Cadmium is associated with many clinical problems, including
heart failure, malignancies, cataract development and lung-related issues. The toxicity symptoms
of arsenic, mercury and lead, depend on the form in which these are consumed. Arsenic promotes
protein coagulation with coenzymes and inhibits adenosine triphosphate (ATP) synthesis. Arsenic
oxides are carcinogenic and serve violent mortality in higher concentrations. Arsenic is associated
with several disorders, such as Guillain-Barre syndrome, which is known as an anti-immune
disorder. In this condition, the patient’s Peripheral Nervous System (PNS) is attacked by the human
immune system, which results in nerve inflammation and muscle weakness.
Mercury (Hg) is found widely in higher aquatic plant species in several forms, including HgS,
Hg2+, methyl-Hg and Hg0 (Wang and Greger 2004, Kamal et al. 2004, Malar et al. 2015). Higher
quantities of Hg2+ are known to be highly phytotoxic to plant cells, causing apparent lesions and
physiological problems, which affects the closing of stomata in leaves and blockage of transportation
of water in plants (Zhou et al. 2007). It is also observed that high amounts of Hg2+ have been
found to interfere with the plant mitochondrial function, resulting in the alteration of cell membrane
components, including lipids, and affecting the cellular metabolism (Messer et al. 2005, Cargnelutti
et al. 2006).
Lead impacts negatively on plant shape, growth, photosynthesis and metabolic activities by
interfering with essential enzymes, limiting seed germination (Zulfiqar et al. 2019). Another problem
caused by increased Pb concentrations is oxidative stress. In plants, it promotes the formation of
Reactive Oxygen Species (ROS) (Reddy et al. 2005). Plants which act as accumulators are highly
resistant to heavy metals in their habitat and can tolerate high concentrations. These plants can use
a different mechanism, such as Exclusion, which limits the excessive transportation of metals in
the plant. A constant concentration of different metals can be maintained in the shoots. Secondly,
Inclusion, a mechanism that prompts the plants to absorb the maximum metals from the soil. Thirdly,
Bioaccumulation, which is the build-up of the toxic metal concentrations in the plants. Table 10.2
summarizes the different toxic metals and their impacts on plants.
nutrients is an important factor for microbial growth that affects pollutant degradation. Several
microbial mediated approaches are used. Biostimulation is the technique that involves a selective
enrichment of existing soil microbe for efficient bioremediation; bioaugmentation involves the
addition of a microbial strain for biodegradation of specific contaminant; bioaccumulation consists
of the uptake of pollutants in the microbial cells for storage and utilization in metabolic activity
(Mehana et al. 2020). Biofilm biosorption involves the removal of contaminants via surface
adsorption. Bacterial bioremediation has the benefit of being resistant to the presence of particular
pollutants.
10.4.1.1 Bioleaching
The bioleaching approach by acidophilic bacteria facilitates the solubility of heavy metals from
sediments and contaminated sites. It predominantly affects the concentrations of toxic metals,
including iron and sulfur, in sediments. Thus, groups of bacteria oxidize iron and sulfur compounds
include: Firmicutes (e.g., Alicyclobacillus sp., Sulfobacillus sp.), Nitrospirae (e.g., Leptospirillum sp.),
Proteobacteria (e.g., Acidithiobacillus sp., Acidiphilium sp., Acidiferrobacter sp., Ferrovum sp.),
Actinobacteria, archaea (Crenarchaeota). These microorganisms are known as bioleaching
microorganisms (Akcil et al. 2015). They can establish an acidic environment by oxidizing minerals,
and therefore dissolve toxic metals ions into an aqueous medium (Akcil et al. 2015).
Chemolitho-autotrophic bacteria Leptospirillum oxidized Fe+2 under acidic conditions
(pH = 1.5 – 1.8) in a sample adulterated with uranium (U92) (Bertrand et al. 2015), whereas
Thiobacillus is found to be capable of generating energy by oxidation of sulfur and thiosulfate (Xin
et al. 2009). Fungal strain Aspergillus niger is also found to be efficient for leaching the elements
and removing pollutants from sediment (Zeng et al. 2015).
The efficiency of bioleaching depends on various parameters, including abiotic stresses,
characteristics of ions present in sediments, pH, and size of sediments. The biotic factors like the
presence of microbial communities, metabolic pathways used for the process, and adaptability to
minerals also play a significant role. As and Fe are the most leached metals from sediments. With
these, a small percentage of Cu is also leachable. The bioleaching process is not suitable for Hg.
The remediating capabilities of indigenous heterotrophic bacterial isolates of Bacillus have been
observed in the leaching of toxic metals from polluted sites (Štyriaková et al. 2016).
These methods are viable both financially and environmentally. According to different research
studies, bioleaching of toxic metals is more effective than the conventional leaching methods
(Liu et al. 2003, Deng et al. 2013). However, several significant disadvantages restrict the utilization
of this technique for an effective detoxification process. Excessive metal concentrations usually limit
the development or functioning of susceptible microbes (Collinet and Morin 1990). A high load of
solid materials and organic materials in polluted sources also hinders microorganisms, resulting in
decreased bioleaching effectiveness (Cho et al. 2002).
10.4.1.2 Biosurfactant
Biosurfactants are synthesized by various classes of microorganisms, including fungi, and are
known to be a substitute for conventional leaching owing to factors such as low toxicity, high
biodegradability and significant environmental friendliness (Shekhar et al. 2015). Biosurfactants
are the molecules that serve as biological chelating agents for a number of heavy metals. The
biosurfactant-producing microorganisms are used at heavy metal-contaminated locations (Pacwa
Płociniczak et al. 2011). Biosurfactants are composed of different functional groups, including fatty
acids, glycolipids and phospholipids (Pacwa-Płociniczak et al. 2011). Rhamnolipid is the most
studied biosurfactant for heavy metal elimination (Mulligan et al. 1999).
Z-glycolipid in Burkholderia sp. is capable of being utilized as a biosurfactant to remove a
mixture of toxic metals like Pb, As, Cu, Cd, Zn and As from polluted soil. Due to large acid-soluble
properties and high chelation with biosurfactants, Mn, Zn and Cd have been more efficiently
removed from the soil. The biosurfactant, rhamnolipid, demonstrated efficient heavy metal removal
from polluted river sediments in exchangeable, carbonate bound or iron-manganese oxide–bound
fractions (Chen et al. 2017).
Using biosurfactants to remove pollutants from soil, waste and sediments is an extremely
interesting approach. It may also be used for the pre-treatment of contaminated areas before
proceeding with stabilization/solidification, natural attenuation or electrokinetic processes. This
method has the edge over other bioprocesses due to its ability to operate effectively even at high pH
up to 11, while other bioprocesses (such as bioleaching and biosorption) work effectively only at low
pH levels. Biosurfactants with excellent degradability and biocompatibility provide a considerable
advantage with high environmental acceptability (Mulligan 2005). However, the exorbitantly high
Metal(loid)s Toxicity and Bacteria Mediated Bioremediation 175
cost of biosurfactant production and the poor productivity of biosurfactants serve as a barrier to their
commercial applications in bioremediation. Further, endeavors are needed in this area to recycle and
reuse biosurfactants in order to decrease total costs via the development of cost-effective recovery
techniques.
10.4.1.3 Bioaccumulation
Bioaccumulation is a complicated and dynamic procedure that requires the absorption and
deposition of heavy metal ions in microbial intracellular components. The metabolism-dependent
heavy metal transporting mechanisms determine its effectiveness. Heavy metals move across cell
membranes via ion channels, protein pathways and carrier-mediated movement (Mishra and Malik
2013). The lipid bilayer allows for passive diffusion and could even enable toxic metals to enter the
cell through endocytosis. Heavy metal active transportation is observed in a number of microbial
species, including pseudomonas, Micrococcus, Bacillus and Aspergillus. However, this technique
is not very useful for bioremediation since heavy metal deposited in the cell may mediate harmful
effects on microbial metabolism. The efflux pathway releases heavy metals into the environment
after a certain level of build up. Despite these restrictions, bioaccumulation has been utilized as
a complete bioremediation method when the requirements of a minimum growing medium with
low harmful impact on cells during treatment are required. In one study, the removal of Hg via
bioaccumulation was shown to be more efficient than biosorption when using a live seaweed
Ulva lactuca (Henriques et al. 2015).
10.4.1.4 Biosorption
Microorganisms, including bacteria, are used to extract toxic metals from sites like contaminated
sewage, soil and sediments (Bano et al. 2018). Biosorption is a metabolism-independent, passive
absorption mechanism that binds heavy metals to the cell membrane. It comprises both the
physical and chemical bonding such as electrostatic, covalent, exopolysaccharides, ion-exchange,
Van der Waal’s force and microprecipitation with different functional groups (Montazer-Rahmati
et al. 2011). Several factors, including acidity, temperature, ionic strength of the environment,
permeability, origin, pre-treatment of biosorbents, amount and speciation of heavy metals, affect the
adsorption process (Zhu et al. 2013, Fomina and Gadd 2014). The bacteria Ochrobactrum MT180101
showed high biosorption efficiency for copper chelated with other compounds (Sun et al. 2021).
The affinity for metal binding is attributed to the presence of functional groups on the microbial
cell that includes alkanes, amides, amines, as well as negatively charged exopolysaccharides
(EPS). Furthermore, EPS modifications like acetylation, carboxymethylation, and methylation may
increase the affinity for metal ions (Gupta and Sar 2020).
10.4.1.5 Bioprecipitation
The process of altering soluble toxic metals ions into insoluble species like hydroxides, carbonates,
phosphates and sulfide groups by microorganisms is known as bioprecipitation. In this approach,
microbe-assisted precipitation is not contingent on the activity of microbes, as it could be present
in live and dead cells. Moreover, it may result in the precipitation of toxic metals that are directly
linked to the cells. Ambient factors such as pH levels and redox potentials influence the efficiency of
bioprecipitation. Sulfate-reducing bacteria produce hydrogen sulfide in an anaerobic environment
using organic substances as a nucleophile, and they can precipitate metal ions. Biological
oxidation of soluble ferrous iron under oxidative circumstances produces Fe (III) hydroxides that
co-precipitate additional ions such as sulfide, Cd, and U indirectly (Kaplan et al. 2016, Rinklebe and
Shaheen 2017). The leaching capability and the availability of toxic metals in sediments are greatly
reduced by these insoluble complexes. By oxidation, the sulfide group is resolubilized into the
aqueous phase. Hence, it is a critical process to track fluctuations in environmental redox potential
and microorganisms’ activity.
176 Bioremediation for Sustainable Environmental Cleanup
10.4.1.6 Biotransformation
Microbes interact with toxic metals and alter their form to a reasonably less toxic form via multiple
biological reactions such as condensation, hydrolysis, formation of new carbon bonds, isomerization
and introduction of functional groups, etc. (Table 10.2) (Guo et al. 2019). In Bacillus species
converts Cr (VI) to Cr (III) under physical conditions like pH (7–9), temperature (between 30ºC
and 40ºC), and Cr (VI) levels (50–250 mg L–1) (Lei et al. 2019). Under optimal conditions of pH 7
and at 37ºC, this species completely reduced the Cr (VI) concentration of 120 mg L–1 in just 48 hr.
By using acetate as a carbon source, a heterogenous anaerobic colony containing Anaerolineaceae,
Spirochaeta and Spirochaetaceae exhibited the capability to the reduction of Cr (VI) and V (V)
with high efficiency of 97 and 99.1%, sequentially (Wu et al. 2019). In the presence of hematite and
dissolved organic compounds, Geobacter sulfurreducens showed the capacity to eliminate Cr (VI).
Microorganisms use a sequence of methylation reactions for the conversion of As to volatile
forms. Microorganisms transform arsenic trioxide (As2O3) into volatile toxic tri-methyl arsine
(CH3)3As and increase the mobility and release of As into the environment. Acinetobacter sp.
and Micrococcus sp. are likewise found to be converting poisonous As (III) into nontoxic and less
soluble As (III) and lessen its toxicity. In methylation of Hg, bacteria convert mercuric ions into
methyl mercuric, which increases the bioavailability of Hg through food sources in the aquatic
environment. The biochemical processes of methylation of Hg and SO42– in contaminated sites are
significantly associated with reducing the pollutant levels in sediment pore water (Hines et al. 2012).
elevated doses of dangerous chemicals and pollutants better than planktonic cells (Davey and
O’toole 2000, Matz and Kjelleberg 2005).
Biofilm-producing bacteria are efficient to bioremediate as they are imprisoned in an EPS
that also immobilizes pollutants during breakdown (Mah and O’Toole 2001). Due to the low
concentration of oxygen towards the center, all the microbes, including aerobes and anaerobes,
heterotrophs (nitrifying bacteria) and sulfate reducers are present in close proximity in the three-
dimensional network of EPS, facilitating quicker degradation of various contaminants in natural and
artificial systems (Sutherland 2001). Toxic metals are removed from the aqueous environment using
EPS from cyanobacteria as a biosorbent (de Philippis et al. 2011). Several enzymes are present in
biofilms EPS, which detoxify heavy metals and organic compounds. Due to the presence of many
negatively charged functional groups in EPS, it acts as a trap for metals and metalloids, allowing
the formation of chelates with toxic metals and organic contaminants and thus facilitating their
elimination (Li and Yu 2014). Different metals like zinc, lead, nickel, magnesium, cadmium, iron,
manganese and copper are known to bind to EPS (Pal and Paul 2008). Nutrient restriction may
lead to enhancement in EPS and copper production and allow to absorb more pollutants from the
polluted site. The EPS of phosphorus-accumulating bacteria in biofilms serves as a reservoir and
aids in the bioremediation and recovery of phosphorus from wastewater (Zhang et al. 2013).
Biofilms may be observed in natural environments such as soil, plants, sediments, streams,
ponds, rivers and man-made environments. Fungus, bacteria, protozoa and algae make up biofilms
that form on the water’s surface.
During bloom periods, fragile structures termed flocs are formed. The floc-activity of activated
sludge plants is used to treat municipal wastewater. Slow-sand filters are used to separate organic
chemicals and metals from the natural environment by using biofilms (Burmølle et al. 2006).
Planctomycetes found in seaweed biofilms in coastal habitats can be used to remove nitrogen from
wastewater by anammox reactions by converting ammonia to dinitrogen anaerobically (Kartal et al.
2010).
Biofilms are extensively used for risk assessment or as an indication for monitoring and assessing
toxic metal pollution. There may be a change in biofilm structural and physiological functions in the
presence of toxic compounds. Microbial biofilms could accumulate contaminants, and contaminants
offer an easy attachment site for biofilms. They are applied as an indicator system (Bengtsson and
Øvreås 2010). Microbes are the first organisms in aquatic habitats to interact with minerals and
pollutants, and as a result, biofilms may be used as a risk assessment tool in aqueous bodies (Fuchs
et al. 1997). Several biofilm marker properties, including biomass production, microbial species,
photosynthetic machinery, pigment formation and enzyme activity, may be used in monitoring
environmental pollution. The microbial species present in river biofilms vary depending on the
season and as well as the level of pollution (Peacock et al. 2004). Contaminated sites with various
toxic metals such as zinc and cadmium may also affect the microbial species; hence, it could be
related to the level of contamination (Brümmer et al. 2000). Biofilm sampling has been shown
to provide a more accurate assessment of heavy metal pollution in aquatic microbial populations
(Bricheux et al. 2013). The algal biofilm could be used as an indicator of pollution due to changes
in biomass in the presence of heavy metals (Ancion et al. 2013). The pigment-formation property
of the biofilm may be altered after exposure to hazardous chemicals, which serves as a biomarker
(Navarro et al. 2002, Dorigo et al. 2004, Dewez et al. 2008).
Acidothiobacillus ferrooxidans, for instance, has demonstrated the ability to achieve industrial-scale
bioleaching (Zhang et al. 2018). Acidocella aromatica and Acidiphilium symbioticum demonstrated
the capability to reduce vanadium ions and biosorption of cadmium cations under extreme pH levels
(Okibe et al. 2016).
A more effective method for toxic metal removal could be accomplished by using a microbe’s
consortium. The bioremediation of these pollutants by using an acid-isolated acidophilic microbe
consortia was performed on a contaminated sediment site. An acidophilic microbial composed of
Acidothiobacillus thiooxidans, Leptospirillum ferrooxidans and Acidiphilium cryptum demonstrated
effectiveness in extracting over 90% of Cu2+, Cd2+, Hg2+ and Zn2+ (Beolchini et al. 2009)
(Table 10.3).
Halophilic bacteria provide significant benefits in the remediation of hazardous contaminants in
extreme saline environments. As marine bacteria can survive at high salinity, bioremediation using
marine bacteria may be a viable alternative for cleaning seawater composed of toxic metals. Vibrio
harveyi has shown a significant potential to acquire cadmium ions. The adsorption capacity was
observed to be up to 23.3 mg Cd2+/g dry cells (Abd-Elnaby et al. 2011). The other marine bacteria,
Enterobacter cloaceae, could make complexes with Cd, Cu and Co from mixed-salts solutions
(Iyer et al. 2005). In conjunction with marine bacteria, thermophilic microbes have high biosorption
capability, indicating these microbes have a high potential for removing contaminants from polluted
environments (Özdemir et al. 2013).
Another approach is to study new extremophilic bacterial enzymes: extremozymes, which
have peculiar structure-function properties such as stability at high temperatures, severe pH, high
ionic strength, in the presence of organic solvents and toxic metals (Cabrera and Blamey 2018).
Extremophilic bacteria such as Metallosphaera sedula, Leptospirillum ferriphilum and Sulfolobus
solfataricus have been sequenced, and segments containing the Hg-resistance gene merA have been
discovered.
Table 10.3. Efficiency of extremophiles in bioremediation of toxic metals.
As a result, extremophiles could be used in the elimination of toxic metals from toxic locations
and sludges. However, more research into advancing technology for investigating microbial
surroundings and gaining insight into the pathways analysis which influence microbial activity and
metal degradation metabolic pathways under severe environments is necessary.
of open biotechnological applications. The most pressing issue is developing genetically modified
bacteria with an acceptable degree of environmental assurance for field release in bioremediation.
The efforts to evaluate the performance of modified bacteria under severe environmental conditions
include endurance and horizontal gene transfer capacity, that may influence the native microflora
of the environment. To avoid this, field bacteria are specially developed for in vitro bioremediation.
It is unclear if the intentional discharge of genetically engineered bacteria for bioremediation has
any negative impact on native microorganisms (Sayler and Ripp 2000). As a result, the survival of
genetically modified microorganisms in a hostile environment remains a major concern.
10.8 Conclusions
Several studies have been conducted for the bioremediation of toxic metals from the environment.
Microbial mediated bioremediation is found to be a sustainable approach. However, certain
areas need to be more focused on and explored so that they can be applied for commercial use.
As bioremediation in most studies is successful in the laboratory and controlled conditions, hence
there is a need to explore the mechanism to carry out bioremediation under natural environment
without controlled conditions. Besides this, GMOs are being used due to their high capability to
perform bioremediation in comparison to wild-type microbial strains. However, there are several
risks associated with this technology, such as ecological disturbance and horizontal gene transfer,
which cause the limited use of GMOs in bioremediation. Extremophiles need to be explored more,
and there might be unculturable microbes present in the environment, hence the need to develop the
process to culture these unexplored microbes. The process development for scale-up is required for
more commercial applications.
Acknowledgments
This work is supported by Grant no. BT/RLF/Re-entry/ 40/ 2017 from the Department of Biotechnology
(DBT), Ministry of Science and Technology and SERB project file no: EEQ/2020/000614 Govt. of
India. We acknowledge the CSIR pool scientific scheme Grant no. 9103, Govt. of India. Pictures are
created by using www.biorender.com.
References
Abdel-Gadir, A., R. Berber, J. B. Porter, P. D. Quinn, D. Suri, P. Kellman et al. 2016. Detection of metallic cobalt
and chromium liver deposition following failed hip replacement using T2* and R2 magnetic resonance. J.
Cardiovasc. Magn. Reson. 18: 29.
Abd-Elnaby, H., G. M. Abou-Elela and N. A. El-Sersy. 2011. Cadmium resisting bacteria in Alexandria Eastern
Harbor (Egypt) and optimization of cadmium bioaccumulation by Vibrio harveyi. Afr. J. Biotech. 10:
3412–3423.
Abedin, Md. J., J. Cotter-Howells and A. A. Meharg. 2002. Arsenic uptake and accumulation in rice (Oryza sativa L.)
irrigated with contaminated water. Plant Soil. 240: 311–319.
Akcil, A., C. Erust, S. Ozdemiroglu, V. Fonti and F. Beolchini. 2015. A review of approaches and techniques used in
aquatic contaminated sediments: metal removal and stabilization by chemical and biotechnological processes.
J. Clean. Prod. 86: 24–36.
Al Khateeb, W. and H. Al-Qwasemeh. 2014. Cadmium, copper and zinc toxicity effects on growth, proline content
and genetic stability of Solanum nigrum L., a crop wild relative for tomato; comparative study. Physiol. Mol.
Biol. Plants. 20: 31–39.
Al Osman, M., F. Yang and I. Y. Massey. 2019. Exposure routes and health effects of heavy metals on children.
Biomet.: Int. J. Role. Met. Ion. Biol. Biochem. Med. 32: 563–573.
Ancion, P. Y., G. Lear, A. Dopheide and G. D. Lewis. 2013. Metal concentrations in stream biofilm and sediments and
their potential to explain biofilm microbial community structure. Environ. Pollut. 173: 117–124.
Andreoli, V. and F. Sprovieri. 2017. Genetic aspects of susceptibility to mercury toxicity: an overview. Int. J. Environ.
Res. Public Health. 14: 93.
Arya, S. K. and B. K. Roy. 2011. Manganese induced changes in growth, chlorophyll content and antioxidants activity
in seedlings of broad bean (Vicia faba L.). J. Environ. Biol. 32: 707–711.
Metal(loid)s Toxicity and Bacteria Mediated Bioremediation 181
Asrar, Z., R. Khavari-Nejad and H. Heidari. 2005. Excess manganese effects on pigments of Mentha spicata at
flowering stage. Arch. Agron. Soil Sci. 51: 101–107.
Ay, T. A., iran, O. O. Fawole, S. O. Adewoye and M. A. Ogundiran. 2009. Bioconcentration of metals in the body
muscle and gut of Clarias gariepinus exposed to sublethal concentrations of soap and detergent effluent. J.
Cell Anim. Biol. 3: 113–118.
Aycicek, M., O. Kaplan and M. Yaman. 2008. Effect of cadmium on germination, seedling growth and metal contents
of sunflower (Helianthus annus L.). Asian J. Chem. 20: 2663–2672.
Azad, M. A. K., L. Amin and N. M. Sidik. 2014. Genetically engineered organisms for bioremediation of pollutants
in contaminated sites. Chin. Sci. Bull. 59: 703–714.
Baby, J., J. S. Raj, E. T. Biby, P. Sankarganesh, M. V. Jeevitha, S. U. Ajisha et al. 2010. Toxic effect of heavy metals
on aquatic environment. Int. J. Biol. Chem. Sci. 4.
Bano, A., J. Hussain, A. Akbar, K. Mehmood, M. Anwar, M. S. Hasni et al. 2018. Biosorption of heavy metals by
obligate halophilic fungi. Chemosphere. 199: 218–222.
Barta, J. A., C. A. Powell and J. P. Wisnivesky. 2019. Global epidemiology of lung cancer. Annals Glob. Health.
85(1): 8.
Belda, E., R. G. A. van Heck, M. José Lopez-Sanchez, S. Cruveiller, V. Barbe, C. Fraser et al. 2016. The revisited
genome of Pseudomonas putida KT2440 enlightens its value as a robust metabolic chassis. Environ. Microbiol.
18: 3403–3424.
Bengtsson, M. M. and L. Øvreås. 2010. Planctomycetes dominate biofilms on surfaces of the kelp Laminaria
hyperborea. BMC Microbiol. 10: 261.
Beolchini, F., A. Dell’Anno, L. de Propris, S. Ubaldini, F. Cerrone and R. Danovaro. 2009. Auto- and heterotrophic
acidophilic bacteria enhance the bioremediation efficiency of sediments contaminated by heavy metals.
Chemosphere. 74: 1321–1326.
Bertrand, J. C., P. Caumette, P. Lebaron, R. Matheron, P. Normand and T. Sime-Ngando. 2015. Environmental
microbiology: fundamentals and applications. Springer Netherlands.
Bondarenko, O., T. Rõlova, A. Kahru and A. Ivask. 2008. Bioavailability of Cd, Zn and Hg in soil to nine recombinant
luminescent metal sensor bacteria. Sensors. 8(11): 6899–6923.
Bricheux, G., G. le Moal, C. Hennequin, G. Coffe, F. Donnadieu, C. Portelli et al. 2013. Characterization and evolution
of natural aquatic biofilm communities exposed in vitro to herbicides. Ecotoxicol. Environ. Saf. 88: 126–134.
Brümmer, I. H. M., W. Fehr and I. Wagner-Döbler. 2000. Biofilm community structure in polluted rivers: abundance
of dominant phylogenetic groups over a complete annual cycle. Appl. Environ. Microbiol. 66: 3078–3082.
Brutti, C. S., R. R. Bonamigo, T. Cappelletti, G. M. Martins-Costa and A. P. S. Menegat. 2013. Occupational and
non-occupational allergic contact dermatitis and quality of life: a prospective study. Anal. Bras. de Derma.
88: 670–671.
Burmølle, M., J. S. Webb, D. Rao, L. H. Hansen, S. J. Sørensen and S. Kjelleberg. 2006. Enhanced biofilm formation
and increased resistance to antimicrobial agents and bacterial invasion are caused by synergistic interactions
in multispecies biofilms. Appl. Environ. Microbiol. 72: 3916–3923.
Cabrera, M. Á. and J. M. Blamey. 2018. Biotechnological applications of archaeal enzymes from extreme
environments. Biol. Res. 51: 1–15.
Cargnelutti, D., L. A. Tabaldi, R. M. Spanevello, G. de Oliveira Jucoski, V. Battisti, M. Redin et al. 2006. Mercury
toxicity induces oxidative stress in growing cucumber seedlings. Chemosphere. 65: 999–1006.
Chakravarty, R. and P. C. Banerjee. 2012. Mechanism of cadmium binding on the cell wall of an acidophilic bacterium.
Bioresour. Technol. 108: 176–183.
Chen, P., L. Yan, F. Leng, W. Nan, X. Yue, Y. Zheng et al. 2011. Bioleaching of realgar by Acidithiobacillus
ferrooxidans using ferrous iron and elemental sulfur as the sole and mixed energy sources. Bioresour. Technol.
102: 3260–3267.
Chen, W., Y. Qu, Z. Xu, F. He, Z. Chen, S. Huang et al. 2017. Heavy metal (Cu, Cd, Pb, Cr) washing from river
sediment using biosurfactant rhamnolipid. Environ. Sci. Pollut. Res. 24: 16344–16350.
Cheng, S. 2003. Heavy metal pollution in China: origin, pattern and control. Environ. Sci. Pollut. Res. 10: 192–198.
Cho, K. S., H. W. Ryu, I. S. Lee and H. M. Choi. 2002. Effect of solids concentration on bacterial leaching of heavy
metals from sewage sludge. J. Air Waste Manag. Association. 52: 237–243.
Collinet, M. N. and D. Morin. 1990. Characterization of arsenopyrite oxidizing Thiobacillus. Tolerance to arsenite,
arsenate, ferrous and ferric iron. Antonie van Leeuwenhoek. 57: 237–244.
Dangi, A. K., K. K. Dubey and P. Shukla. 2017. Strategies to Improve Saccharomyces cerevisiae: technological
advancements and evolutionary engineering. Indian J. Microbiol. 57: 378–386.
Das, K. K., S. N. Das and S. A. Dhundasi. 2008. Nickel, its adverse health effects & oxidative stress. Indian J. Med.
Res. 128: 412.
182 Bioremediation for Sustainable Environmental Cleanup
Davey, M. E. and G. A. O’toole. 2000. Microbial biofilms: from ecology to molecular genetics. Microbiol. Mol. Biol.
Rev. 64: 847–867.
Demirevska-Kepova, K., L. Simova-Stoilova, Z. Stoyanova, R. Hölzer and U. Feller. 2004. Biochemical changes in
barley plants after excessive supply of copper and manganese. Environ. Exp. Bot. 52: 253–266.
Deng, X., L. Chai, Z. Yang, C. Tang, Y. Wang and Y. Shi. 2013. Bioleaching mechanism of heavy metals in the
mixture of contaminated soil and slag by using indigenous Penicillium chrysogenum strain F1. J. Hazard.
Mater. 248-249: 107–114.
de Philippis, R., G. Colica and E. Micheletti. 2011. Exopolysaccharide-producing cyanobacteria in heavy metal
removal from water: molecular basis and practical applicability of the biosorption process. Appl. Microbiol.
Biotechnol. 92: 697–708.
Dewez, D., O. Didur, J. Vincent-Héroux and R. Popovic. 2008. Validation of photosynthetic-fluorescence parameters
as biomarkers for isoproturon toxic effect on alga Scenedesmus obliquus. Environ. Pollut. 151: 93–100.
Doncheva, S., K. Georgieva, V. Vassileva, Z. Stoyanova, N. Popov and G. Ignatov. 2005. Effects of succinate on
manganese toxicity in pea plants. J. Plant. Nutr. 28: 47–62.
Dorigo, U., X. Bourrain, A. Bérard and C. Leboulanger. 2004. Seasonal changes in the sensitivity of river microalgae
to atrazine and isoproturon along a contamination gradient. Sci. Total. Environ. 318: 101–114.
D’Souza, C. and R. Peretiatko. 2002. The nexus between industrialization and environment: a case study of Indian
enterprises. Environ. Manag. H. 13: 80–97.
Du, X., Y.-G. Zhu, W.-J. Liu and X.-S. Zhao. 2005. Uptake of mercury (Hg) by seedlings of rice (Oryza sativa L.)
grown in solution culture and interactions with arsenate uptake. Environ. Exp. Bot. 54: 1–7.
Edwards, K. J., P. L. Bond, T. M. Gihring and J. F. Banfield. 2000. An archaeal iron-oxidizing extreme acidophile
important in acid mine drainage. Science. 287: 1796–1799.
Flemming, H. C. and J. Wingender. 2001. Relevance of microbial extracellular polymeric substances (EPSs) - Part I:
Structural and ecological aspects. Water Sci. Technol. IWA Publishing. 43(6): 1–8.
Fomina, M. and G. M. Gadd. 2014. Biosorption: current perspectives on concept, definition and application.
Bioresour. Technol. 160: 3–14.
Fuchs, S., T. Haritopoulou, M. Schäfer and M. Wilhelmi. 1997. Heavy metals in freshwater ecosystems introduced by
urban rainwater runoff—Monitoring of suspended solids, river sediments and biofilms. Water. Sci. Technol.
36: 277–282.
García-Esquinas, E., M. Pollán, J. G. Umans, K. A. Francesconi, W. Goessler, E. Guallar et al. 2013. Arsenic exposure
and cancer mortality in a US-based prospective cohort: the strong heart study. Caner Epidemiol. Pren. Biomark.
22: 1944–1953.
Gieg, L. M., S. J. Fowler and C. Berdugo-Clavijo. 2014. Syntrophic biodegradation of hydrocarbon contaminants.
Curr. Opin. Biotechnol. 27: 21–29.
Gumulya, Y., N. J. Boxall, H. N. Khaleque, V. Santala, R. P. Carlson and A. H. Kaksonen. 2018. In a quest for
engineering acidophiles for biomining applications: challenges and opportunities. Genes. 9: 116.
Guo, T., L. Li, W. Zhai, B. Xu, X. Yin, Y. He et al. 2019. Distribution of arsenic and its biotransformation genes in
sediments from the East China Sea. Environ. Pollut. 253: 949–958.
Gupta, A. and P. Sar. 2020. Characterization and application of an anaerobic, iron and sulfate reducing bacterial
culture in enhanced bioremediation of acid mine drainage impacted soil. J. Environ. Sci. Health. - Part A Toxic/
Hazard. Subst. Environ. Eng. 55: 464–482.
Hamada, A. J., S. C. Esteves and A. Agarwal. 2013. A comprehensive review of genetics and genetic testing in
azoospermia. Clinics. 68: 39–60.
Heck, J. E., A. S. Park, J. Qiu, M. Cockburn and B. Ritz. 2014. Risk of leukemia in relation to exposure to ambient air
toxics in pregnancy and early childhood. Int. J. Hyg. Environ. Health. 217: 662–668.
Henriques, B., L. S. Rocha, C. B. Lopes, P. Figueira, R. J. R. Monteiro, A. C. Duarte et al. 2015. Study on
bioaccumulation and biosorption of mercury by living marine macroalgae: prospecting for a new remediation
biotechnology applied to saline waters. Chem. Eng. J. 281: 759–770.
Hines, M. E., E. N. Poitras, S. Covelli, J. Faganeli, A. Emili, S. Žižek et al. 2012. Mercury methylation and
demethylation in Hg-contaminated lagoon sediments (Marano and Grado Lagoon, Italy). Estuar. Coast Shelf
Sci. 113: 85–95.
Holliger, C. and A. J. B. Zehnder. 1996. Anaerobic biodegradation of hydrocarbons. Curr. Opin. Biotechnol. 7:
326–330.
Horemans, B., P. Breugelmans, J. Hofkens, E. Smolders and D. Springael. 2013. Environmental dissolved organic
matter governs biofilm formation and subsequent linuron degradation activity of a linuron-degrading bacterial
consortium. Appl. Environ. Microbiol. 79: 4534–4542.
Ivask, A., H. C. Dubourguier, L. Põllumaa and A. Kahru. 2011. Bioavailability of Cd in 110 polluted topsoils to
recombinant bioluminescent sensor bacteria: effect of soil particulate matter. J. Soils Sediments. 11: 231–237.
Metal(loid)s Toxicity and Bacteria Mediated Bioremediation 183
Iyer, A., K. Mody and B. Jha. 2005. Biosorption of heavy metals by a marine bacterium. Mar. Pollut. Bull.
50: 340–343.
Jan, A. T., A. Ali and Q. M. R. Haq. 2011. Glutathione as an antioxidant in inorganic mercury induced nephrotoxicity.
J. Postgrad. Med. 57: 72.
Jameson, E., O. F. Rowe, K. B. Hallberg and D. B. Johnson. 2010. Sulfidogenesis and selective precipitation of
metals at low pH mediated by Acidithiobacillus spp. and acidophilic sulfate-reducing bacteria. Hydrometall.
104: 488–493.
Jayakumar, K., C. A. Jaleel and P. Vijayarengan. 2007. Changes in growth, biochemical constituents, and antioxidant
potentials in radish (Raphanus sativus L.) under Cobalt Stress. Turk. J. Biol. 31: 127–136.
Jayakumar, K., C. Jaleel and M. Azooz. 2008. Phytochemical changes in green gram (Vigna radiata) under cobalt
stress. Glob. J. Mol. Sci. 3(2): 46–49.
Jia, Y., H. Huang, M. Zhong, F. H. Wang, L. M. Zhang and Y. G. Zhu. 2013. Microbial arsenic methylation in soil and
rice rhizosphere. Environ. Sci. Technol. 47: 3141–3148.
Jung, J. H., N. Y. Choi and S. Y. Lee. 2013. Biofilm formation and exopolysaccharide (EPS) production by Cronobacter
sakazakii depending on environmental conditions. Food. Microbiol. 34: 70–80.
Kamal, M., A. E. Ghaly, N. Mahmoud and R. Côté. 2004. Phytoaccumulation of heavy metals by aquatic plants.
Environ. Int. 29: 1029–1039.
Kaplan, D. I., R. Kukkadapu, J. C. Seaman, B. W. Arey, A. C. Dohnalkova, S. Buettner et al. 2016. Iron mineralogy
and uranium-binding environment in the rhizosphere of a wetland soil. Sci. Total Environ. 569-570: 53–64.
Kartal, B., J. G. Kuenen and M. C. M. van Loosdrecht. 2010. Sewage treatment with anammox. Science. 328:
702–703.
Kashefi, K. and D. R. Lovley. 2000. Reduction of Fe(III), Mn(IV), and toxic metals at 100°C by Pyrobaculum
islandicum. Appl. Environ. Microbiol. 66: 1050–1056.
Kennady, V., R. Verma and V. Chaudhiry. 2018. Detrimental impacts of heavy metals on animal reproduction: a
review. J. Entomol. Zool. Stud. 06: 27–30.
Kodama, H., C. Fujisawa and W. Bhadhprasit. 2012. Inherited copper transport disorders: biochemical mechanisms,
diagnosis, and treatment. Curr. Drug Metab. 13: 237–250.
Kostal, J., R. Yang, C. H. Wu, A. Mulchandani and W. Chen. 2004. Enhanced arsenic accumulation in engineered
bacterial cells expressing ArsR. Appl. Environ. Microbiol. 70: 4582–4587.
Lacal, J., J. A. Reyes-Darias, C. García-Fontana, J. L. Ramos and T. Krell. 2013. Tactic responses to pollutants and
their potential to increase biodegradation efficiency. J. Appl. Microbiol. 114: 923–933.
Lei, P., H. Zhong, D. Duan and K. Pan. 2019. A review on mercury biogeochemistry in mangrove sediments: hotspots
of methylmercury production? Sci. Total. Environ. 680: 140–150.
Leyssens, L., B. Vinck, C. Van Der Straeten, F. Wuyts and L. Maes. 2017. Cobalt toxicity in humans—a review of the
potential sources and systemic health effects. Toxicol. 387: 43–56.
Li, W. W. and H. Q. Yu. 2014. Insight into the roles of microbial extracellular polymer substances in metal biosorption.
Bioresour. Technol. 160: 15–23.
Liu, H. L., C. W. Chiu and Y. C. Cheng. 2003. The effects of metabolites from the indigenous Acidithiobacillus
thiooxidans and temperature on the bioleaching of cadmium from soil. Biotechnol. Bioeng. 83: 638–645.
Liu, S., Y. Zheng, Y. Ma, A. Sarwar, X. Zhao, T. Luo et al. 2019. Evaluation and proteomic analysis of lead adsorption
by lactic acid bacteria. Int. J. Mol. Sci. 20: 5540.
Lim, H. W., S. A. B. Collins, J. S. Resneck, J. L. Bolognia, J. A. Hodge, T. A. Rohrer et al. 2017. The burden of skin
disease in the United States. J. Am. Acad. Dermatol. 76: 958–972.
Lin, H.-J., T.-I. Sung, C.-Y. Chen and H.-R. Guo. 2013. Arsenic levels in drinking water and mortality of liver cancer
in Taiwan. J. Hazard. Mater. 262: 1132–1138.
Mah, T. F. C. and G. A. O’Toole. 2001. Mechanisms of biofilm resistance to antimicrobial agents. Trends Microbiol.
9: 34–39.
Mahurpawar, M. 2015. Effects of heavy metals on human health. Int. J. Res. -Granth. 3: 1–7.
Malar, S., S. V. Sahi, P. J. C. Favas and P. Venkatachalam. 2015. Assessment of mercury heavy metal toxicity-
induced physiochemical and molecular changes in Sesbania grandiflora L. Int. J. Environ. Sci. Technol.
12: 3273–3282.
Matz, C. and S. Kjelleberg. 2005. Off the hook—How bacteria survive protozoan grazing. Trends. Microbiol.
13: 302–307.
Meeker, J. D., M. G. Rossano, B. Protas, V. Padmanahban, M. P. Diamond, E. Puscheck et al. 2010. Environmental
exposure to metals and male reproductive hormones: circulating testosterone is inversely associated with
blood molybdenum. Fertil. Steril. 93: 130–140.
184 Bioremediation for Sustainable Environmental Cleanup
Mehana, E.-S. E., A. F. Khafaga, S. S. Elblehi, M. E. Abd El-Hack, M. A. E. Naiel, M. Bin-Jumah, S. I. Othman and
A. A. Allam. 2020. Biomonitoring of heavy metal pollution using Acanthocephalans parasite in ecosystem:
An updated overview. Animals 10(5): 811.
Meharg, A. A. 2004. Arsenic in rice—understanding a new disaster for South-East Asia. Trends. Plant Sci. 9: 415–417.
Messer, R. L. W., P. E. Lockwood, W. Y. Tseng, K. Edwards, M. Shaw, G. B. Caughman et al. 2005. Mercury (II) alters
mitochondrial activity of monocytes at sublethal doses via oxidative stress mechanisms. J. Biomed. Mater.
Res. Part B: Appl. Biomater. 75: 257–263.
Miqueleto, A. P., C. C. Dolosic, E. Pozzi, E. Foresti and M. Zaiat. 2010. Influence of carbon sources and C/N ratio on
EPS production in anaerobic sequencing batch biofilm reactors for wastewater treatment. Bioresour. Technol.
101: 1324–1330.
Mishra, A. and A. Malik. 2013. Recent advances in microbial metal bioaccumulation. Crit. Rev. Environ. Sci. Technol.
43: 1162–1222.
Montazer-Rahmati, M. M., P. Rabbani, A. Abdolali and A. R. Keshtkar. 2011. Kinetics and equilibrium studies on
biosorption of cadmium, lead, and nickel ions from aqueous solutions by intact and chemically modified
brown algae. J. Hazard. Mater. 185: 401–407.
More, T. T., J. S. S. Yadav, S. Yan, R. D. Tyagi and R. Y. Surampalli. 2014. Extracellular polymeric substances of
bacteria and their potential environmental applications. J. Environ. Manag. 144: 1–25.
Mujtaba Munir, M. A., G. Liu, B. Yousaf, M. U. Ali, Q. Abbas and H. Ullah. 2020. Synergistic effects of biochar and
processed fly ash on bioavailability, transformation and accumulation of heavy metals by maize (Zea mays L.)
in coal-mining contaminated soil. Chemosphere. 240: 124845.
Mulligan, C. N., R. N. Yong, B. F. Gibbs, S. James and H. P.J. Bennett. 1999. Metal removal from contaminated soil
and sediments by the biosurfactant surfactin. Environ. Sci. Technol. 33: 3812–3820.
Mulligan, C. N. 2005. Environmental applications for biosurfactants. Environ. Pollut. 133: 183–198.
Musilova, J., J. Arvay, A. Vollmannova, T. Toth and J. Tomas. 2016. Environmental contamination by heavy metals in
region with previous mining activity. Bull. Environ. Contam. Toxicol. 97: 569–575.
Nagajyoti, P. C., K. D. Lee and T. V. M. Sreekanth. 2010. Heavy metals, occurrence and toxicity for plants: a review.
Environ. Chem. Lett. 8: 199–216.
Navarro, E., H. Guasch and S. Sabater. 2002. Use of microbenthic algal communities in ecotoxicological tests for the
assessment of water quality: the Ter river case study. J. Appl. Phycol. 14: 41–48.
Nogawa, K., E. Kobayashi, Y. Okubo and Y. Suwazono. 2004. Environmental cadmium exposure, adverse effects and
preventive measures in Japan. Biometals. 17: 581–587.
Nordberg, G. F. 2004. Cadmium and health in the 21st century—historical remarks and trends for the future.
Biometals. 17: 485–489.
Okereafor, U., M. Makhatha, L. Mekuto, N. Uche-Okereafor, T. Sebola and V. Mavumengwana. 2020. Toxic metal
implications on agricultural soils, plants, animals, aquatic life and human health. Int. J. Environ. Res. Public
Health. 17: 2204.
Okibe, N., M. Maki, D. Nakayama and K. Sasaki. 2016. Microbial recovery of vanadium by the acidophilic bacterium,
Acidocella aromatica. Biotechnol. Lett. 38: 1475–1481.
Özdemir, S., E. Klnç, A. Poli and B. Nicolaus. 2013. Biosorption of heavy metals (Cd2+, Cu2+, Co 2+, and Mn2+) by
thermophilic bacteria, Geobacillus thermantarcticus and Anoxybacillus amylolyticus: equilibrium and kinetic
studies. Bioremediat. J. 17: 86–96.
Pacwa-Płociniczak, M., G. A. Płaza, Z. Piotrowska-Seget and S. S. Cameotra. 2011. Environmental applications of
biosurfactants: recent advances. Int. J. Mol. Sci. 12: 633–654.
Pal, A. and A. K. Paul. 2008. Microbial extracellular polymeric substances: central elements in heavy metal
bioremediation. Indian J. Microbio. 48: 49–64.
Patel, J., Q. Zhang, R. M. L. McKay, R. Vincent and Z. Xu. 2010. Genetic engineering of caulobacter crescentus for
removal of cadmium from water. Appl. Biochem. Biotechnol. 160: 232–243.
Peacock, A. D., Y. J. Chang, J. D. Istok, L. Krumholz, R. Geyer, B. Kinsall et al. 2004. Utilization of microbial
biofilms as monitors of bioremediation. Microb. Ecol. 47: 284–292.
Plum, L. M., L. Rink and H. Haase. 2010. The essential toxin: impact of zinc on human health. Int. J. Environ. Res.
Public Health. 7: 1342–1365.
Pratt, L. A. and R. Kolter. 1999. Genetic analyses of bacterial biofilm formation. Curr. Opin. Microbiol. 2: 598–603.
Rahman, S. F., R. S. Kantor, R. Huddy, B. C. Thomas, A. W. van Zyl, S. T. L. Harrison et al. 2017. Genome-resolved
metagenomics of a bioremediation system for degradation of thiocyanate in mine water containing suspended
solid tailings. Microbiol. Open. 6: e00446.
Rajaganapathy, V., F. Xavier, D. Sreekumar and P. K. Mandal. 2011. Heavy metal contamination in soil, water and
fodder and their presence in livestock and products: a review. J. Environ. Sci. Technol. 4: 234–249.
Ray, R. R. 2016. Adverse hematological effects of hexavalent chromium: an overview. Interdiscip. Toxicol. 9: 55–65.
Metal(loid)s Toxicity and Bacteria Mediated Bioremediation 185
Ray, R. R. 2017. Review article. Adverse hematological effects of hexavalent chromium: an overview. Interdiscip.
Toxicol. 9: 55–65.
Reddy, A. M., S. G. Kumar, G. Jyothsnakumari, S. Thimmanaik and C. Sudhakar. 2005. Lead induced changes
in antioxidant metabolism of horsegram (Macrotyloma uniflorum (Lam.) Verdc.) and bengalgram (Cicer
arietinum L.). Chemosphere. 60: 97–104.
Rinklebe, J. and S. M. Shaheen. 2017. Redox chemistry of nickel in soils and sedimentsA review. Chemosphere.
179: 265–278.
Ripp, S., D. E. Nivens, Y. Ahn, C. Werner, J. Jarrell IV, J. P. Easter et al. 2000. Controlled field release of a
bioluminescent genetically engineered microorganism for bioremediation process monitoring and control.
Environ. Sci. Technol. 34: 846–853.
Rojas, L. A., C. Yáñez, M. González, S. Lobos, K. Smalla and M. Seeger. 2011. Characterization of the metabolically
modified heavy metal-resistant Cupriavidus metallidurans strain MSR33 generated for mercury bioremediation.
PLoS ONE. 6(3): e17555.
Romero-González, M., B. C. Nwaobi, J. M. Hufton and D. J. Gilmour. 2016. Ex-situ bioremediation of U(VI) from
contaminated mine water using Acidithiobacillus ferrooxidans Strains. Front. Environ. Sci. 4: 39.
Romero-Puertas, M. C., M. Rodríguez-Serrano, F. J. Corpas, M. Gómez, L. a. D. Río and L. M. Sandalio. 2004.
Cadmium-induced subcellular accumulation of O2·− and H2O2 in pea leaves. Plant Cell Environ.
27: 1122–1134.
Saavedra, A., P. Aguirre and J. C. Gentina. 2020. Biooxidation of Iron by Acidithiobacillus ferrooxidans in the
presence of D-Galactose: understanding its influence on the production of EPS and cell tolerance to high
concentrations of iron. Front. Microbiol. 11: 759.
Sayler, G. S. and S. Ripp. 2000. Field applications of genetically engineered microorganisms for bioremediation
processes. Curr. Opin. Biotechnol. 11: 286–289.
Sharma, B. and P. Shukla. 2020. Futuristic avenues of metabolic engineering techniques in bioremediation. Biotech.
Appl. Biochem. https://doi.org/10.1002/bab.2080.
Sheikh, I. 2016. Cobalt Poisoning: a comprehensive review of the literature. J. Med. Toxicol. Clin. Forensic Med. 2.
Shekar, C. C., D. Sammaiah, T. Shasthree and K. J. Reddy. 2011. Effect of mercury on tomato growth and yield
attributes. Int. J. Pharm. Bio. Sci. 2.
Shekhar, S., A. Sundaramanickam and T. Balasubramanian. 2015. Biosurfactant producing microbes and their
potential applications: a review. Crit. Rev. Environ. Sci. Technol. 45: 1522–1554.
Sheldon, A. and N. W. Menzies. 2005. The effect of copper toxicity on growth and morphology of Rhodes grass
(Chlorisgayana) in solution culture. Plant Soil, 278.
Sobha, K., A. Poornima, P. Harini and K. Veeraiah. 2007. A study on biochemical changes in the fresh water fish,
Catla catla (Hamilton) exposed to the heavy metal toxicant cadmium chloride. Kathmandu Univ. J. Sci. Eng.
Technol. 3: 1–11.
Spangler, A. H. and J. G. Spangler. 2009. Groundwater manganese and infant mortality rate by county in North
Carolina: an ecological analysis. EcoHealth. 6: 596–600.
Štyriaková, I., I. Štyriak, A. Balestrazzi, C. Calvio, M. Faè and D. Štyriaková. 2016. Metal leaching and reductive
dissolution of iron from contaminated soil and sediment samples by indigenous bacteria and Bacillus isolates.
Soi. Sediment Contam. 25: 519–535.
Sun, W., B. Zhu, F. Yang, M. Dai, S. Sehar, C. Peng et al. 2021. Optimization of biosurfactant production from
Pseudomonas sp. CQ2 and its application for remediation of heavy metal contaminated soil. Chemosphere
265: 129090.
Sutherland, I. W. 2001. The biofilm matrix—An immobilized but dynamic microbial environment. Trends Microbiol.
9: 222–227.
Tahri, N., W. Bahafid, H. Sayel and N. el Ghachtouli. 2013. Biodegradation: involved microorganisms and genetically
engineered microorganism. Ch. 11. pp. 289–320. In: R. Chamy and F. Rosenkranz [eds.]. Biodegradation -
Life of Science. IntechOpen.
Vieira, C., S. Morais, S. Ramos, C. Delerue-Matos and M. B. P. P. Oliveira. 2011. Mercury, cadmium, lead and arsenic
levels in three pelagic fish species from the Atlantic Ocean: Intra- and inter-specific variability and human
health risks for consumption. Food Chem. Toxicol. 49: 923–932.
Wang, Y. and M. Greger. 2004. Clonal differences in mercury tolerance, accumulation, and distribution in willow. J.
Environ. Qual. 33: 1779–1785.
Wang, M., J. Zou, X. Duan, W. Jiang and D. Liu. 2007. Cadmium accumulation and its effects on metal uptake in
maize (Zea mays L.). Bioresour. Technol. 98: 82–88.
Woo, S., S. Yum, H.-S. Park, T.-K. Lee and J.-C. Ryu. 2009. Effects of heavy metals on antioxidants and stress-
responsive gene expression in Javanese medaka (Oryzias javanicus). Comp. Biochem. Physiol. Part C:
Toxicol. Pharmacol. 149: 289–299.
186 Bioremediation for Sustainable Environmental Cleanup
Wu, C. H., T. K. Wood, A. Mulchandani and W. Chen. 2006. Engineering plant-microbe symbiosis for rhizoremediation
of heavy metals. Appl. Environ. Microbiol. 72: 1129–1134.
Wu, M., Y. Li, J. Li, Y. Wang, H. Xu and Y. Zhao. 2019. Bioreduction of hexavalent chromium using a novel strain
CRB-7 immobilized on multiple materials. J. Hazard. Mater. 368: 412–420.
Xie, X., W. Zhu, N. Liu and J. Liu. 2013. Bacterial community composition in reclaimed and unreclaimed tailings of
Dexing copper mine, China. Afr. J. Biotechnol. 12(30): 4841–4849.
Xin, B., D. Zhang, X. Zhang, Y. Xia, F. Wu, S. Chen et al. 2009. Bioleaching mechanism of Co and Li from spent
lithium-ion battery by the mixed culture of acidophilic sulfur-oxidizing and iron-oxidizing bacteria. Bioresour.
Technol. 100: 6163–6169.
Yang, S., M. Yu and J. Chen. 2017. Draft genome analysis of Dietzia sp. 111N12-1, isolated from the South China Sea
with bioremediation activity. Braz. J. Microbiol. 48: 393–394.
Yao, J., L. Tian, Y. Wang, A. Djah, F. Wang, H. Chen et al. 2008. Microcalorimetric study the toxic effect of hexavalent
chromium on microbial activity of Wuhan brown sandy soil: An in vitro approach. Ecotoxicol. Environ. Saf.
69: 289–295.
Yourtchi, M. S. and H. R. Bayat. 2013. Effect of cadmium toxicity on growth, cadmium accumulation and
macronutrient content of durum wheat (Dena CV.). Int. J. Agric. Crop Sci. (IJACS). 6: 1099–1103.
Yusuf, M., Q. Fariduddin, P. Varshney and A. Ahmad. 2012. Salicylic acid minimizes nickel and/or salinity-induced
toxicity in Indian mustard (Brassica juncea) through an improved antioxidant system. Environ. Sci. Pollut.
Res. 19: 8–18.
Zdrojewicz, Z., E. Popowicz and J. Winiarski. 2016. Nickel—role in human organism and toxic effects. Pol. Merk.
Lek. 41: 115–118.
Zeng, X., S. Wei, L. Sun, D. A. Jacques, J. Tang, M. Lian et al. 2015. Bioleaching of heavy metals from contaminated
sediments by the Aspergillus niger strain SY1. J. Soil Sediment 15: 1029–1038.
Zhang, H. L., W. Fang, Y. P. Wang, G. P. Sheng, R. J. Zeng, W. W. Li et al. 2013. Phosphorus removal in an enhanced
biological phosphorus removal process: roles of extracellular polymeric substances. Environ. Sci. Technol.
47: 11482–11489.
Zhang, S., L. Yan, W. Xing, P. Chen, Y. Zhang and W. Wang. 2018. Acidithiobacillus ferrooxidans and its potential
application. Extremophiles. 22: 563–579.
Zheng, X., L. Chen, M. Chen, J. Chen and X. Li. 2019. Functional metagenomics to mine soil microbiome for novel
cadmium resistance genetic determinants. Pedosphere. 29: 298–310.
Zhou, Z. S., S. Q. Huang, K. Guo, S. K. Mehta, P. C. Zhang and Z. M. Yang. 2007. Metabolic adaptations to mercury-
induced oxidative stress in roots of Medicago sativa L. J. Inorg. Biochem. 101: 1–9.
Zhu, X., H. Yu, H. Jia, Q. Wu, J. Liu and X. Li. 2013. Solid phase extraction of trace copper in water samples
via modified corn silk as a novel biosorbent with detection by flame atomic absorption spectrometry. Anal.
Methods. 5: 4460–4466.
Zulfiqar, U., M. Farooq, S. Hussain, M. Maqsood, M. Hussain, M. Ishfaq, M. Ahmad and M. Z. Anjum. 2019. Lead
toxicity in plants: impacts and remediation. J. Environ. Manag. 250: 109557.
Chapter 11
Lead Induced Toxicity,
Detoxification and
Bioremediation
Shalini Dhiman,1 Arun Dev Singh,1 Isha Madaan,2,7
Raman Tikoria,3 Driti Kapoor,4 Priyanka Sharma,5
Nitika Kapoor,6 Geetika Sirhindi,2 Puja Ohri 3 and
Renu Bhardwaj1,*
11.1 Introduction
Lead is a hazardous element that comes from various technogenic sources like ammunition, batteries-
based industries, bangle manufacturing, building material, ceramic ware, cosmetics, gasoline,
glassware, plastic pipes, paints with Pb pigments, petrochemicals, radiation protection, leaded fuels,
bullets, fishing sinkers, mining, smelting, electroplating of the metallic ores (Dotaniya et al. 2020).
All the possible routes of Pb exposure beyond its permissible limits lead to Pb toxicity in soil,
water, air, humans, animals as well as in plant systems. Pb inside soil disrupts soil properties and
soil ecosystem mainly by reducing nutrients availability to plants, which in turn effects soil-forming
processes, microbiota, soil health, crop quality and productivity. However, Pb uptake beyond
permissible limits inside plant systems causes many abnormal morphological and physiological
symptoms like disturbance in photosynthesis, water potential, nutrient uptake, respiration and causes
nuclear damage. Moreover, plants also show tolerance and detoxification mechanism against heavy
metal stress such as compartmentalization, metallothionine, phytochelatins, Pb-immobilization by
plants roots exudates, organic acid, etc. (Mitra et al. 2020).
1
Department of Botanical and Environmental Sciences, Guru Nanak Dev University, Amritsar, Punjab, India, 143005.
2
Department of Botany, Punjabi University, Patiala, India, 147002.
3
Department of Zoology, Guru Nanak Dev University, Amritsar, India, 143005.
4
Department of Botany, School of Bioengineering and Biosciences, Lovely Professional University, Phagwara, India,
144402.
5
School of Bioengineering Sciences & Research, MIT-ADT Loni Kalbhor, Pune, Maharashtra, India, 412201.
6
PG Department of Botany, Hans Raj Mahila Maha Vidyalaya, Jalandhar, Punjab, India, 144623.
7
Government College of Education, Jalandhar, Punjab, India, 144001.
* Corresponding author: [email protected]
188 Bioremediation for Sustainable Environmental Cleanup
Harmless removal of Pb from Pb-contaminated sites is the need of this modern century. Such
safe disposal and proper remediation of Pb from polluted sites have been done using various modern
tools and techniques such as bioremediation, especially phytoremediation. Some plants are very
capable of remediating contaminants from contaminated sites, and such processes are collectively
also known as phytoremediation (Jagetiya and Kumar 2020). Soils polluted with Pb are remediated
by other living forms such as fungi, algae, microbes, etc. A considerable number of studies have
been successfully carried out on Pb-bioremediation because these techniques are acceptable, safer,
cheaper and more ecofriendly than any other techniques for the removal of Pb.
therefore less accessible for plant uptake (Punamiya et al. 2010). An investigatory report given by
Kumpiene et al. (2017) showed that lead is a versatile element present in the soil with a background
amount of 27 mg kg–1. Lead-contaminated soil is divided into five categories depending on the level
of contamination, i.e., extremely low (< 150 ppm), low (150–400 ppm), moderate (400–1000 ppm),
high (1000–2000 ppm) and extremely high (> 2000 ppm). The degree of Pb contamination also
varies from one season to another, and in different mediums (Patel et al. 2010). Reports suggested
that a high level of lead is reported in some places where anthropogenic activities are prominent. For
example—at smelting sites, lead concentration in soil ranged from 10 to 7100 mg kg–1 (Chlopecka
et al. 1996), road dust (105–110 mg kg–1) (Bi et al. 2018), mining site (132–45016 mg kg–1) (Higueras
et al. 2017). Further, the industrial site contained lead at a level of 42–131 mg kg–1 (He et al. 2017);
in previous garden soil, it was 1020–1030 mg kg–1 (Egendorf et al. 2018). Similarly, Pb in flooded
soil was in between 105–115 mg kg–1 (Antić-Mladenović et al. 2017), and in the soil of shooting
ranges was 32,500–33,500 mg kg–1 (Mariussen et al. 2018).
metals as compared to edible crops. Further, crops grown on land continuously supplied by sewage
water also have a high level of lead, which ultimately leads to a bad color of the crops (Bhupal Raj
et al. 2009). Onion crops quality was found inferior in the soil irrigated by lead-contaminated water
in Ratlam, India (Meena et al. 2020).
Severe loss of yield in crops was observed in soybean, gram, fenugreek and garlic. These crops
got affected by Pb contaminated water, which decreased farmers’ output (Panwar et al. 2010).
11.2.1.2.3 Industries
Agricultural soil/lands that are situated closer to industrial areas are more prone to lead contamination
because of leakage of lead-containing pollutants (Saha et al. 2013). Pb level was found greater in
the soil surrounding the Coimbatore area near an electroplating and paint industry. Similarly, in
Tamil Nadu’s Dindigul district, the cement factory is the major source of lead poisoning (Meena
et al. 2020). Lead contamination into the soil is also caused by industries that make daily life
products such as Pb-based paints, solder, ceramics and pesticides. Industries that are involved in
mining smelting and tailing release high amounts of lead into nearby soil (Mitra et al. 2020, Anju
and Banerjee 2011).
11.2.1.2.4 Urbanization
Lead increase in urban soils might be due to increasing industrialization. Combustion of leaded
petrol results in vehicular emission containing tetraethyl Pb, which contributes significantly to lead
in urban areas. As per USEPA standards, the threshold limit of Pb in the soil is 400 mg/L, and in
portable water, it is 0.01 mg/L (Bureau of Indian Standard) (Mitra et al. 2020).
Lead Induced Toxicity, Detoxification and Bioremediation 191
DISTRESSED RESPIRATION
IMPAIRED PHOTOSYNTHESIS
Decreased chlorophyll content Downregulation of TCA cycle enzymes
Altered chloroplast structure Improper electron and proton transport
Inefficient Hill reaction and
& Calvin
Calvincycle
cycle
Pb in Soil
Figure 11.1. Uptake of lead from soil and its toxic effects within the plant.
Pb in PLANT SYSTEM
Detoxification Strategies
within cytoplasm before it enters vacuoles and chloroplast, thereby minimizing the toxic effects of
Pb on the plant cell (Pourrut et al. 2013).
Lead Bioremediation
Plants Mediated
Microbes Mediated
Phytoremediation Processes
Rhizosphere
Mechanistic Approach by Microbes Phytoextraction
Phytostabilization,
Biosorption, Bioaccumulation Bioaccumulation of ions
Biomineralization via oxidation state Rhizofiltration
transformation and enzymes Phtovolatization
Efflux of ions
Extracellular and intracellular sequestration
Siderophores
Metal chelation
Figure 11.3. Overview of bioremediation strategies to reduce lead toxicity.
by the functional groups present/occurring on their surface to adsorb metal ions (Yin et al. 2016).
In soil, bacteria restricted Pb concentration via fostering insoluble lead complexes with hydroxide,
sulfide and carbonates, transforming the active form into stable insoluble state.
Further, sequestration of toxic ions is mainly through exopolysaccharide, an organic
polysaccharide with smaller proteins and lipids. Many microorganisms like Xanthomonas,
Bacillus, Agrobacterium, Alcaligenes, Pseudomonas spp., etc., have been identified as genera of
EPS-producing organisms to achieve heavy metal remediation by utilizing the charged property of
EPS, where they are incorporated with abundant anionic functional groups (Tayang and Songachan
2021). This mechanism is critical in the process of biomineralization, metal ions biosorption and
bioaccumulation (Thakare et al. 2021). Similarly, Chen et al. (2015) described Bacillus thuringiensis
as a potential biosorbent for Pb (II) transformation. Further, Bacillus cereus could transform Pb into
Pb hydroxyapatite via enzymatic action. An experiment conducted by scientists revealed that the
bacterium Streptomyces and Staphylococcus showed a prominent binding affinity for lead and other
metals. Therefore, they can be effectively used for the biosorption of lead (Sahmoune 2018).
Li et al. (2017) findings revealed that the bacterial strains of Pseudomonas sps., can efficiently
absorb Pb (II) from wastewater sites. For the first time, Borremans et al. (2001) found a lead-
resistance strain, i.e., CH34 in R. metallidurans, which enhances uptake and efflux mechanism by
the pbr operon system. Later many studies have shown the involvement of specific genes expression
for resistance to Pb by metallothionein proteins, specifically in P. aeruginosa (Kumari and Das
2019). Kang et al. (2016) confirmed microbial (bacterial) remediation of Pb-contaminated soils due
to the function of precipitation, sequestration or variation in the oxidation state of Pb. They revealed
the synergistic effect of bacterial consortium (E. cloacae, Sporosarcina, Viridibacillus arenosi, and
Enterobacter cloacae) on a mixture of Pb along with other heavy metals against single strain culture.
These bacteria are accountable for the transformation of HMs by enzymes production (Huang et al.
2009). It was observed that Bacillus iodinium, Klebsiella aerogenes and Bacillus pumilus precipitate
Pb (II) into PbS 9 (Govarthanan et al. 2013).
It has been noted that c‐type cytochromes and porin–cytochrome proteins in outer membrane
proteins in the microbes are involved in declining contaminants toxicity (Shi et al. 2016). Several
studies have shown that microbes transformed the state of metal by changing the valence status of
metals via redox-mediated processes (Dixit et al. 2015, Shi et al. 2016). The bacterial organisms
such as Bacillus sps., A. eutrophus, Pseudomonas sps., produce siderophores enabling extraction of
Pb from soil (Naik and Dubey 2017, Kalita and Joshi 2017). Similarly, in another report, the positive
interaction among siderophores and metal Pb and Ni was revealed by the bacteria P. aeruginosa
(Braud et al. 2009, 2010). The Pteris vittata plant exhibited rapid growth in the Pb-contaminated
area, which was enabled due to Pseudomonas spp., resistance against metal Pb via the process of
extracellular sequestration (Manzoor et al. 2019). Therefore, these examples of evidence show the
significance of siderophore-producing bacteria, which cause extraction and mobilization of Pb from
contaminated soil.
Table 11.1. Role of various plant species for the remediation of Pb.
Different strategies adopted by microorganisms to survive in heavy metal contaminated soils are
extrusion of metal ions by using metal efflux pumps, biotransformation of ions, intra/extracellular
metal sequestration, enzymatic usage, exopolysaccharide (EPS) generation, and metallothionein
and bio surfactants synthesis, etc. (Dixit et al. 2015, Igiri et al. 2018). Microorganisms can
further detoxify the metal ions by several different methods, including ion exchange, electrostatic
interaction, precipitation, surface complexation, etc. (Yang et al. 2015). Microorganisms have
negatively charged groups on their cell surface that facilitate them to bind to cationic metal ions
(Gavrilescu 2004).
Fungal hyphae remediate the heavy metal contaminated soils by intracellular sequestration
of toxic metal ions. Chitin, lipids, mineral ions, N-polysaccharide, polyphosphates and proteins
are major constituents of the cell wall of fungi. Fungal hyphae and their spores can eradicate the
heavy metal ions from the soil by ATPase pump-mediated uptake, extracellular and intracellular
precipitation and change in the oxidation state of metal ions. On the fungi cell wall outer surface,
there is the presence of various metal ions binding ligands/functional groups that enhanced the rate
of binding of toxic metals to hyphae, thereby reducing the availability of these toxic metals to plants.
The foremost metal-binding ligands present on the surface of the cell wall are the hydroxyl group,
carboxyl group, phosphoryl group, sulfate group, sulfite group, ester, amine group, carboxylate
group and sulfanyl group. Out of these functional groups, the amine group is most involved in metal
absorption as it can bind with both cationic as well as anionic metal ions by surface complexation
and electrostatic interaction, respectively (Gupta et al. 2015, Xie et al. 2016).
198 Bioremediation for Sustainable Environmental Cleanup
Various reports exhibited that active as well as inactive fungal cells, play an essential role in the
adsorption of inorganic metal ions (Srivastava and Thakur 2006, Tiwari et al. 2013, Igiri et al. 2018).
Lakkireddy and Kües (2017) reported that Coprinopsis atramentaria could accumulate 94.7% of
800 mg L−1 of Pb2+. As a result, it is being identified as a good heavy metal ion accumulator for
mycoremediation. Some of the fungal biomasses viz. Aspergillus niger, Rhizopus oryzae, Penicillium
chrysogenum and Saccharomyces cerevisiae are also effective in converting the most hazardous
oxidation state of heavy metal to less toxic/non-toxic oxidation state of heavy metals (Park et al.
2005). Inoculum of Arbuscular Mycorrhizal Fungi (AMF) is found to enhance the Pb remediation
efficiency of Japanese clover (Kummerowia striata (Thunb.)) and barnyard grass (Echinochloa
crus-galli L.) (Chen et al. 2005). Fungi like Aspergillus sp. and Coprinopsis sp. are largely used
as biosorbents for eliminating toxic metals with great potential for metal absorption and recovery
(Akar et al. 2005, Dursun et al. 2003).
Biosurfactants produced by fungi also play a major role in cleaning heavy metals from
contaminated soil. Luna et al. (2016) reported that anionic biosurfactant from Candida sphaerica
has 79% removal efficiency for Pb from heavy metal contaminated soil. The biosurfactant was
found to be effective in removing the exchangeable, oxide, carbonate and organic fractions of heavy
metals by forming complexes with metal ions.
Yeast biosurfactants are also found to be effective in cleaning heavy metals and petroleum
derivatives from contaminated soils by reducing soil permeability. The crude biosurfactant
significantly reduced the concentration of Pb and other heavy metals from the test sample of soil.
Biosurfactants being amphoteric in nature, not only help in eliminating heavy metals but can also
be applied to remove hydrophobic organic compounds. They reduce the interfacial tension and
solubilize hydrocarbons in the aqueous phase or capture the oil droplets within their micelles. On
the other hand, anionic nature biosurfactants capture the metal ions through electrostatic interactions
or complexation (Rufino et al. 2011). In recent years, biosurfactants have received a lot of interest
for their biodegradable nature, low toxicity and diversity. Several yeast strains such as S. cerevisiae,
Rhodotorula pilimanae, Hansenula polymorpha, Yarrowia lipolytica and Rhodotorula mucilage
have been utilized to convert more toxic forms of heavy metals to non-toxic ones (Ksheminska
et al. 2008, Chatterjee et al. 2012).
Phycoremediation is an important aspect that deals with removing or degrading heavy metals
from contaminated sites with the help of algal biomasses. Features that make dead algae biomass
an ideal candidate for the removal of heavy metals include the presence of sulfate and carboxylic
acid functional groups on the cell wall that facilitate metal adsorption and large surface area/volume
ratios. In comparison to other microbial biosorbents, algae are autotrophic, require little nutrients and
produce large amounts of biomass. Heavy metal removal has been achieved using these biosorbents
with a high sorption capacity. Algal biomass bioremediates the heavy metal contaminated effluent
either through adsorption or by integrating inside the cells (Abbas et al. 2014, Chabukdhara et al.
2017, He and Chen 2014).
When compared to other microbial biosorbents, algal biomasses had biosorption effectiveness
of 15.3–84.6%. Ion exchange techniques are used to accomplish this (Mustapha and Halimoon
2015). For successful heavy metals cleanup from the polluted area, algal biomass has been
immobilized using various chemical pretreatments, which lead to the formation of stable cellular
aggregates with appropriate size, efficient mechanical strength, rigidity, porosity and increased
biomass concentration (Laxman and More 2002). Red marine algae Jania ruben L. was found to be
effectively bioadsorb Pb, which was further confirmed by thermal analysis (Hanbali et al. 2014).
Various reports are available in literature regarding the use of algal biomasses in the decontamination
of toxic metals. Goher et al. (2016) reported that dead cells of Chlorella vulgaris could be used for
the removal of copper (Cu2+), cadmium (Cd2+) and lead (Pb2+) ions from an aqueous solution under
several conditions of biosorbent dosage, pH and contact time. The biomass of C. vulgaris removed
cadmium (Cd2+), copper (Cu2+), and lead (Pb2+) at the rate of 95.5, 97.7 and 99.4%, respectively,
constituting a combined solution of 50 mg dm−3 of each metal ion. Thus applying appropriate
Lead Induced Toxicity, Detoxification and Bioremediation 199
microbial inoculum under specific conditions of pH, temperature and dosage could aid plants to
remove heavy metals like Pb from contaminated soil efficiently.
Remediation of toxic contaminants from various substrates by employing a microbial agent,
fungus, is also known as mycoremediation. Fungi have a capacity to recover HMs owing to
filamentous structures that exhibit a charged group on the cell wall. In addition to functional groups,
they also display features of metal transporters by enzymatic activity, vacuolar sequestration and
antioxidant systems (Kumar 2017, Vacar et al. 2021). Fawzy et al. (2017) screened many fungal
species like Emercilla quadrillineata, Rhizopus stolonifier, Aspergillus niger, etc., and observed
an effective resistance even at higher concentrations of Pb in contaminated soil. Similarly, other
fungal isolates such as Rhizophagus irregularis and Funneliformis mosseae showed promising high
biomass of Helianthus annuus against Pb ions application (Hassan et al. 2013). Fungi, particularly
Aspergillus niger, revealed a detoxification mechanism in the environment by the processes of
compartmentalized sequestration, biosorption and chelation with organic acids. Thus, such pathways
can be used for decreasing lead levels through immobilization or mobilization (Bellion et al. 2006,
Iram et al. 2015).
Some species of fungi like Metarhizium and Paecilomyces can convert metallic lead into
chloropyromorphite in a lead mining site. This transformation occurred because of the organic acid
secreted by the fungi, which caused the precipitation of Pb (Rhee et al. 2012, Rigoletto et al. 2020).
Alongwith this, Povedano-Priego et al. (2017) confirmed that biomineralization of lead phosphate
in decaying wood caused tolerance to isolated fungi. They noted that Penicillium and Asperigillus
strains showed more tolerance to heavy metals. A summary of various algae and fungi involved in
the remediation of Pb from contaminated sites is presented in Table 11.2.
11.5 Conclusion
Heavy metal Pb has a large number of applications in industries such as lead used as lead acetate in
sweeteners, installations of drinkable water, paints, additives used in gasoline and many more that
further increase the probability of Pb release, its exposure and penetration in all existing organisms
of our ecosystem. Beyond permissive limits, Pb accumulation inside the organisms creates toxic
hazards and severe morphological and physiological implications and finally decreases the
efficiency of the organism. Thus, the removal of Pb-based hazardous products should be firmly
perused in order to improve and avoid environmental as well as health risks. An essential criteria
for attaining Pb based contaminant-free environment require the development of highly efficient
bioremediation technologies such as phytoremediation, mycoremediation, phycoremediation and
200 Bioremediation for Sustainable Environmental Cleanup
References
Abbas, H. S., M. I. Ismail, M. T. Mostafa and H. A. Sulaymon. 2014. Biosorption of heavy metals: a review. J. Chem.
Sci. Technol. 3: 74–102.
Akar, T., S. Tunali and I. Kiran. 2005. Botrytis cinerea as a new fungal biosorbent for removal of Pb (II) from aqueous
solutions. Biochem. Eng. J. 25(3): 227–235.
Akhtar, N., S. Kha, I. Malook, S. U. Rehman and M. Jamil. 2017. Pb-induced changes in roots of two cultivated rice
cultivars grown in lead-contaminated soil mediated by smoke. Environ. Sci. Pollut. Res. 24: 21298–21310.
Akmal, M. and X. Jianming. 2009. Microbial biomass and bacterial community changes by Pb contamination in
acidic soil. J. Agric. Biol. Sci. 1: 30–37.
Ali, H., E. Khan and M. A. Sajad. 2013. Phytoremediation of heavy metals—concepts and
applications. Chemosphere. 91(7): 869–881.
Aliyu, H. G. and H. M. Adamu. 2014. The potential of maize as phytoremediation tool of heavy metals. Eur. Sci.
J. 10(6).
Amin, H., B. A. Arain, T. M. Jahangir, M. S. Abbasi and F. Amin. 2018. Accumulation and distribution of lead
(Pb) in plant tissues of guar (Cyamopsis tetragonoloba L.) and sesame (Sesamum indicum L.): profitable
phytoremediation with biofuel crops. Geol. Ecol. Landscapes. 2(1): 51–60.
Anju, M. and D. K. Banerjee. 2011. Associations of cadmium, zinc, and lead in soils from a lead and zinc mining area
as studied by single and sequential extractions. Environ. Monit. Assess. 176(1): 67–85.
Antić-Mladenović, S., T. Frohne, M. Kresović, H. J. Stärk, Z. Tomić, V. Ličina and J. Rinklebe. 2017. Biogeochemistry
of Ni and Pb in a periodically flooded arable soil: fractionation and redox-induced (im) mobilization. Environ.
Manag. Today. 186: 141–50.
Aransiola, S. A., U. J. J. Ijah and O. P. Abioye. 2013. Phytoremediation of lead polluted soil by Glycine max L. Appl.
Environ. Soil Sci. 2013.
Arias, J. A., R. Jose, P. Videa, J. T. Ellzey, M. Ren, M. N. Viveros and J. L. Gardea-Torresdey. 2010. Effects of Glomus
deserticola inoculation on prosopis: enhancing chromium and lead uptake and translocation as confirmed by
X-Ray mapping, ICP-OES and TEM techniques. Environ. Exp. Bot. 68(2): 139–148.
Arreghini, S., L. de Cabo, R. Serafini and A. F. de Iorio. 2017. Effect of the combined addition of Zn and Pb on partitioning
in sediments and their accumulation by the emergent macrophyte Schoenoplectus californicus. Environ. Sci.
Pollut. Res. 24(9): 8098–8107.
ATSDR. 2020. “Toxicological Profile for Lead.” Atlanta, Georgia.
Babaeian, E., M. Homaee and R. Rahnemaie. 2016. Chelate-enhanced phytoextraction and phytostabilization of lead-
contaminated soils by carrot (Daucus carota). Arch. Agron. Soil Sci. 62(3): 339–358.
Bahar, M. D., M. M. Megharaj and R. Naidu. 2012. Arsenic bioremediation potential of a new arsenite-oxidizing
bacterium Stenotrophomonas Sp. MM-7 Isolated from Soil. Biodegradat. 23(6): 803–12.
Banat, I. M., A. Franzetti, I. Gandolfi, G. Bestetti, M. G. Martinotti, L. Fracchia, T. J. Smyth and R. Marchant. 2010.
Microbial biosurfactants production, applications and future potential. Appl. Microbiol. Biotechnol. 87(2):
427–44.
Bellion, M., M. Courbot, C. Jacob, D. Blaudez and M. Chalot. 2006. Extracellular and cellular mechanisms sustaining
metal tolerance in ectomycorrhizal fungi. FEMS. Microbiol. Lett. 254(2): 173–181.
Bello, A. O., B. S. Tawabini, A. B. Khalil, C. R. Boland and T. A. Saleh. 2018. Phytoremediation of cadmium-, lead-
and nickel-contaminated water by Phragmites australis in hydroponic systems. Ecol. Eng. 120: 126–133.
Bhupal Raj, G., M. V. Singh, M. C. Patnaik and K. M. Khadke. 2009. Four decades of research on micro and secondary
nutrients and pollutant elements in soils of Andhra Pradesh. Bhopal, India: Indian Institute of Soil Science.
Bi, C., Y. Zhou, Z. Chen, J. Jia and X. Bao. 2018. Heavy metals and lead isotopes in soils, road dust and leafy
vegetables and health risks via vegetable consumption in the industrial areas of Shanghai, China. Sci. Total
Environ. 619-620: 1349–57.
Blaylock, M. J., D. E. Salt, S. Dushenkov, O. Zakharova, C. Gussman, Y. Kapulnik, B. D. Ensley and I. Raskin.
1997. Enhanced accumulation of Pb in Indian mustard by soil-applied chelating agents. Environ. Sci. Technol.
31(3): 860–65.
Borremans, B., J. L. Hobman, A. Provoost, N. L. Brown and D. van der Lelie. 2001. Cloning and functional analysis
of the pbr lead resistance determinant of Ralstonia metallidurans. CH34. J. Bacteriol. 183: 5651–5658.
Lead Induced Toxicity, Detoxification and Bioremediation 201
Bramhachari, P. V., P. B. Kavi Kishor, R. Ramadevi, R. Kumar, B. R. Rao and S. K. Dubey. 2007. Isolation and
characterization of mucous exopolysaccharide produced by Vibrio furnissii VB0S3. J. Microbiol. Biotechnol.
17: 44–51.
Braud, A., K. Jezequel, S. Bazot and T. Lebeau. 2009. Enhanced phytoextraction of an agricultural Cr and Pb
contaminated soil by bioaugmentation with sideropho reproducing bacteria. Chemosphere. 74: 280–286.
Braud, A., V. Geoffroy, F. Hoegy, G. L. A. Mislin and I. J. Schalk. 2010. The siderophores pyoverdine and pyochelin
are involved in Pseudomonas aeruginosa resistence against metals: another biological function of these two
siderophores. Environ. Microbiol. Rep. 2: 419–425.
Camargo, F. A. O., B. C. Okeke, F. M. Bento and W. T. Frankenberger. 2003. In vitro reduction of hexavalent chromium
by a cell-free extract of Bacillus sp. ES 29 stimulated by Cu 2+. Appl. Microbiol. Biotechnol. 62(5): 569–573.
Cecchi, M., C. Dumat, A. Alric, B. Felix-Faure, P. Pradere and M. Guiresse. 2008. Multi-metal contamination of a
calcic cambisol by fallout from a lead-recycling plant. Geoderma. 144(1-2): 287–98.
Chabukdhara, M., S. K. Gupta and M. Gogoi. 2017. Phycoremediation of heavy metals coupled with generation of
bioenergy. In Algal biofuels pp. 163–188. Springer, Cham.
Chatterjee, S., N. C. Chatterjee and S. Dutta. 2012. Bioreduction of chromium (VI) to chromium (III) by a novel yeast
strain Rhodotorula mucilaginosa (MTCC 9315). Afr. J. Biotechnol. 11(83): 14920–14929.
Chen, X., C. Wu, J. Tang and S. Hu. 2005. Arbuscular mycorrhizae enhance metal lead uptake and growth of host
plants under a sand culture experiment. Chemosphere. 60(5): 665–671.
Chen, Z., X. Pan, H. Chen, Z. Lin and X. Guan. 2015. Investigation of lead (II) uptake by Bacillus thuringiensis 016.
World J. Microbiol. Biotechnol. 31: 1729–1736.
Cheng, S. F., C. Y. Huang, K. L. Chen, S. C. Lin and Y. C. Lin. 2016. Phytoattenuation of lead-contaminated agricultural
land using Miscanthus floridulus—An in situ case study. Desalin. Water Treat. 57(17): 7773–7779.
Chibuike, G. U. and S. C. Obiora. 2014. Heavy metal polluted soils: effect on plants and bioremediation methods. Appl.
Environ. Soil Sci. 2014.
Chlopecka, A., J. R. Bacon, M. J. Wilson and J. Kay. 1996. Forms of cadmium, lead, and zinc in contaminated soils
from Southwest Poland. J. Environ. Qual. 25(1): 69–79.
De, J. and N. Ramaiah. 2007. Characterization of marine bacteria highly resistant to mercury exhibiting multiple
resistances to toxic chemicals. Ecol. Indic. 7: 511–520.
Deng, Z., L. Cao, H. Huang, X. Jiang, W. Wang, Y. Shi and R. Zhang. 2011. Characterization of Cd-and Pb-resistant
fungal endophyte Mucor sp. CBRF59 isolated from rapes (Brassica chinensis) in a metal-contaminated soil. J.
Hazard. Mater. 185(2-3): 717–724.
Dixit, R., D. Malaviya, K. Pandiyan, U. B. Singh, A. Sahu, R. Shukla, B. P. Singh, J. P. Rai, P. K. Sharma and H. Lade.
2015. Bioremediation of HMs from soil and aquatic environment: an overview of principles and criteria of
fundamental processes. Sustainability. 7(2): 2189–2212.
Dotaniya, M. K., V. D. Meena, K. Kumar, B. P. Meena, S. L. Jat, M. Lata, A. Ram, C. K. Dotaniya and M. S. Chari.
2016. Impact of biosolids on agriculture and biodiversity: environmental impact on biodiversity. Today and
Tomorrow’s Printer and Publisher, 11–20.
Dotaniya, M. L., V. D. Meena, S. Rajendiran, M. V. Coumar, J. K. Saha, S. Kundu and A. K. Patra. 2017. Geo
accumulation indices of heavy metals in soil and groundwater of Kanpur, India under long term irrigation of
Tannery Effluent. Bull. Environ. Contam. Toxicol. 98(5): 706–11.
Dotaniya, M. L. and J. S. Pipalde. 2018. Soil enzymatic activities as influenced by lead and nickel concentrations in
a vertisol of Central India. Bull. Environ. Contam. Toxicol. 101(3): 380–85.
Dotaniya, M. L., V. D. Meena, J. K. Saha, S. Rajendiran, A. K. Patra, C. K. Dotaniya, H. M. Meena, K. Kumar and
B. P. Meena. 2018a. Environmental impact measurements: tool and techniques. pp. 1–31. In: L. M. T. Martínez,
O. V. Kharissova and B. I. Kharisov [eds.]. Handbook of Ecomaterials. Cham: Springer International
Publishing, Singapore.
Dotaniya, M. L., N. R. Panwar, V. D. Meena, C. K. Dotaniya, K. L. Regar, M. Lata and J. K. Saha. 2018b.
Bioremediation of metal contaminated soil for sustainable crop production. pp. 143–73. In: V. S. Meena [ed.].
Role of Rhizospheric Microbes in Soil. Springer, Singapore.
Dotaniya, M. L., S. Rajendiran, V. D. Meena, M. V. Coumar, J. K. Saha, S. Kundu and A. K. Patra. 2018c. Impact of
long-term application of sewage on soil and crop quality in vertisols of central India. Bull. Environ. Contam.
Toxicol. 101(6): 779–86.
Dotaniya, M. L., C. K. Dotaniya, P. Solanki, V. D. Meena and R. K. Doutaniya. 2020. Lead contamination and its
dynamics in soil–plant system. pp. 83–98. In: D. K. Gupta, S. Chatterjee and C. Walther [eds.]. Lead in Plants
and the Environment. Springer Nature, Switzerland.
Douay, F., A. Pelfrêne, J. Planque, H. Fourrier, A. Richard, H. Roussel and B. Girondelot. 2013. Assessment of
potential health risk for inhabitants living near a former lead smelter. part 1: metal concentrations in soils,
agricultural crops, and homegrown vegetables. Environ. Monit. Assess. 185(5): 3665–80.
202 Bioremediation for Sustainable Environmental Cleanup
Dursun, A. Y., G. Uslu, Y. Cuci and Z. Aksu. 2003. Bioaccumulation of copper (II), lead (II) and chromium (VI) by
growing Aspergillus niger. Process Biochem. 38(12): 1647–1651.
Egendorf, S. P., Z. Cheng, M. Deeb, V. Flores, A. Paltseva, D. Walsh, P. Groffman and H. W. Mielke. 2018. Constructed
soils for mitigating lead (Pb) exposure and promoting urban community gardening: The New York City clean
soil bank pilot study. Landsc. Urban Plan. 175: 184–94.
Farhan, S. N. and A. A. Khadom. 2015. Biosorption of heavy metals from aqueous solutions by Saccharomyces
cerevisiae. Int. J. Ind. Chem. 6(2): 119–130.
Fawzy, E. M., F. F. Abdel-Motaal and S. A. El-Zayat. 2017. Biosorption of heavy metals onto different eco-friendly
substrates. J. Bioremediat. Biodegrad. 8: 394.
Gabr, R. M., S. H. A. Hassan and A. A. M. Shoreit. 2008. Biosorption of lead and nickel by living and non-living cells
of Pseudomonas aeruginosa ASU 6a. Int. Biodeteriorat. Biodegrad. 62: 195–203.
Gavrilescu, M. 2004. Removal of heavy metals from the environment by biosorption. Eng. Life Sci. 4(3): 219–232.
Ghori, N. H., T. Ghori, M. Q. Hayat, S. R. Imadi, A. Gul, V. Altay and M. Ozturk. 2019. Heavy metal stress and
responses in plants. Int. J. Environ. Sci. Technol. 16(3): 1807–1828.
Goher, M. E., A. E. M. AM, A. M. Abdel-Satar, M. H. Ali, A. E. Hussian and A. Napiórkowska-Krzebietke. 2016.
Biosorption of some toxic metals from aqueous solution using non-living algal cells of Chlorella vulgaris. J.
Elementol. 21(3).
Govarthanan, M., K. J. Lee, M. Cho, J. S. Kim, S. Kamala-Kannan and B. T. Oh. 2013. Significance of autochthonous
Bacillus sp. KK1 on biomineralization of lead in mine tailings. Chemosphere. 90: 2267–2272.
Gupta, D. K., H. G. Huang and F. J. Corpas. 2013. Lead tolerance in plants: strategies for phytoremediation. Environ.
Sci. Pollut. Res. 20(4): 2150–2161.
Gupta, V. K. and A. Rastogi. 2008. Biosorption of lead from aqueous solutions by green algae Spirogyra species:
kinetics and equilibrium studies. J. Hazard. Mater. 152(1): 407–414.
Gupta, V. K., A. Nayak and S. Agarwal. 2015. Bioadsorbents for remediation of heavy metals: current status and their
future prospects. Environ. Eng. Res 20(1): 1–18.
Hadi, F. and T. Aziz. 2015. A mini review on lead (Pb) toxicity in Plants. J. Biol. Life Sci. 6(2): 2157–6076.
Hanbali, M., H. Holail and H. Hammud. 2014. Remediation of lead by pretreated red algae: adsorption isotherm,
kinetic, column modeling and simulation studies. Green Chem. Lett. Rev. 7(4): 342–358.
Hasanuzzaman, M., M. H. M. Bhuyan, F. Zulfiqar, A. Raza, S. M. Mohsin, J. A. Mahmud, M. Fujita and V. Fotopoulos.
2020. Reactive oxygen species and antioxidant defense in plants under abiotic stress: revisiting the crucial role
of a universal defense regulator. Antioxidants. 9(8): 681.
Hassan, S. E., M. Hijri and M. St-Arnaud. 2013. Effect of arbuscular mycorrhizal fungi on trace metal uptake
by sunflower plants grown on cadmium contaminated soil. N. Biotechnol. 30(6): 780–7. doi: 10.1016/j.
nbt.2013.07.002s.
He, J. and J. P. Chen. 2014. A comprehensive review on biosorption of heavy metals by algal biomass: materials,
performances, chemistry, and modeling simulation tools. Bioresour. Technol. 160: 67–78.
He, K., Z. Sun, Y. Hu, X. Zeng, Z. Yu and H. Cheng. 2017. Comparison of soil heavy metal pollution caused by
e-waste recycling activities and traditional industrial operations. Environ. Sci. Pollut. Res. 24: 9387–9398.
Higueras, P., J. M. Esbrí, E. García-Ordiales, B. González-Corrochano, M. A. López-Berdonces, E. M. García-
Noguero, J. Alonso-Azcárate and A. Martínez-Coronado. 2017. Potentially harmful elements in soils and
holm-oak trees (Quercus Ilex L.) Growing in mining sites at the Valle de Alcudia Pb-Zn District (Spain)–some
clues on plant metal uptake. J. Geochem. Explor. 182: 166–79.
Huang, S. H., P. Bing, Z. H. Yang, L. Chai and L. Zhou. 2009. Chromium accumulation, microorganism population
and enzyme activities in soils around chromium containing slag heap of steel alloy factory. Trans. Nonferrous
Metals Soc. China. 19: 241–248.
Hynninen, A., T. Touze, L. Pitkanen, D. Mengin-Lecreulx and M. Virta. 2009. An efflux transporter PbrA and a
phosphatase PbrB cooperate in a lead-resistance mechanism in bacteria. Mol. Microbiol. 74: 384–394.
Igiri, B. E., S. I. Okoduwa, G. O. Idoko, E. P. Akabuogu, A. O. Adeyi and I. K. Ejiogu. 2018. Toxicity and
bioremediation of heavy metals contaminated ecosystem from tannery wastewater: a review. J. Toxicol. 2018.
Iqbal, M. and R. G. J. Edyvean. 2004. Biosorption of lead, copper and zinc ions on loofa sponge immobilized biomass
of Phanerochaete chrysosporium. Minerals Eng. 17(2): 217–223.
Iram, S., R. Shabbir, H. Zafar and M. Javaid. 2015. Biosorption and Bioaccumulation of copper and lead by heavy
metal-resistant fungal isolates. Arab. J. Sci. Eng. 40: 1867–1873.
Jagetiya, B. and S. Kumar. 2020. Phytoremediation of lead: a review. Lead Plants Environ., 171–202.
Joshi, P. K., A. Swarup, S. Maheshwari, R. Kumar and N. Singh. 2011. Bioremediation of heavy metals in liquid
media through fungi isolated from contaminated sources. Indian J. Microbiol. 51(4): 482–487.
Lead Induced Toxicity, Detoxification and Bioremediation 203
Jusselme, M. D., F. Poly, E. Miambi, P. Mora, M. Blouin, A. Pando and C. Rouland-Lefèvre. 2012. Effect of
earthworms on plant Lantana Camara Pb-uptake and on bacterial communities in root-adhering soil. Sci.
Total Environ. 416: 200–207.
Kalita, D. and S. R. Joshi. 2017. Study on bioremediation of lead by exopolysaccharide producing metallophilic
bacterium isolated from extreme habitat. Biotechnol. Rep. 16: 48–57.
Kandziora-Ciupa, M., R. Ciepał, A. Nadgórska-Socha and G. Barczyk. 2013. A comparative study of heavy metal
accumulation and antioxidant responses in Vaccinium myrtillus L. leaves in polluted and non-polluted
areas. Environ. Sci. Pollut. Res. 20(7): 4920–4932.
Kang, C. H., Y. J. Kwon and J. S. So. 2016. Bioremediation of heavy metals by using bacterial mixtures. Ecol.
Eng. 89: 64–69. doi: 10.1016/j.ecoleng.2016.01.023.
Khan, I., M. Iqbal, M. Y. Ashraf, M. A. Ashraf and S. Ali. 2016. Organic chelants-mediated enhanced lead (Pb) uptake
and accumulation is associated with higher activity of enzymatic antioxidants in spinach (Spinacea oleracea L.).
J. Hazard. Mater. 317: 352–361.
Ksheminska, H., D. Fedorovych, T. Honchar, M. Ivash and M. Gonchar. 2008. Yeast tolerance to chromium depends
on extracellular chromate reduction and Cr (III) chelation. Food Technol. Biotechnol. 46(4): 419–426.
Kumar, A., H. Touseef, C. Susmita, D. K. Maurya, M. Danish and S. A. Farooqui. 2021. Microbial remediation and
detoxification of heavy metals by plants and microbes. pp. 589–614. In: M. Shah, S. Rodriguez-Couto and
K. Mehta [eds.]. The Future of Effluent Treatment Plants, Elsevier, SBN 9780128229569.
Kumar, V. V. 2017. Mycoremediation: a step toward cleaner environment. pp. 171–187. In: R. Prasad [ed.].
Mycoremediation and Environmental Sustainability. Springer International Publishing: Cham, Switzerland.
Kumari, S. and S. Das. 2019. Expression of metallothionein encoding gene bmtA in biofilm-forming marine bacterium
Pseudomonas aeruginosa N6P6 and understanding its involvement in Pb (II) resistance and bioremediation.
Environ. Sci. Pollut. Res. (28): 28763–28774.
Kumpiene, J., L. Giagnoni, B. Marschner, S. Denys, M. Mench, K. Adriaensen, J. Vangronsveld, M. Puschenreiter
and G. Renella. 2017. Assessment of methods for determining bioavailability of trace elements in soils: a
review. Pedosphere. 27(3): 389–406.
Kunito, T., K. Saeki, K. Nagaoka, H. Oyaizu and S. Matsumoto. 2001. Characterization of copper-resistant bacterial
community in rhizosphere of highly copper-contaminated soil. Eur. J. Soil Biol. 37(2): 95–102.
Kushwaha, A., N. Hans, S. Kumar and R. Rani. 2018. A critical review on speciation, mobilization and toxicity of
lead in soil-microbe-plant system and bioremediation strategies. Ecotoxicol. Environ. Saf. 147: 1035–1045.
Kushwaha, A., R. Rani, S. Kumar, T. Thomas, A. A. David and M. Ahmed. 2017. A new insight to adsorption
and accumulation of high lead concentration by exopolymer and whole cells of lead-resistant bacterium
Acinetobacter junii L. Pb1 isolated from coal mine dump. Environ. Sci. Pollut. Res. 24: 10652–10661.
Lakkireddy, K. and U. Kües. 2017. Bulk isolation of basidiospores from wild mushrooms by electrostatic attraction
with low risk of microbial contaminations. AMB Express. 7(1): 1–22.
Laxman, R. S. and S. More. 2002. Reduction of hexavalent chromium by Streptomyces griseus. Minerals Eng. 15(11):
831–837.
Lee, Y. C. and S. P. Chang. 2011. The biosorption of heavy metals from aqueous solution by Spirogyra and Cladophora
filamentous macroalgae. Bioresour. Technol. 102(9): 5297–5304.
Lenka, S., S. Rajendiran and V. Coumar. 2016. Impact of fertilizers use on environmental quality. In: National Seminar
on Environmental Concern for Fertilizer Use in Future. Krishi Viswavidyalaya, Kalyani.
Li, X., L. Zhang and G. Wang. 2014. Genomic evidence reveals the extreme diversity and wide distribution of the
arsenic-related genes in Burkholderiales. PLoS One. 9(3): e92236.
Li, X., D. Meng, J. Li, H. Yin, H. Liu and X. Liu. 2017. Response of soil microbial communities and microbial
interactions to long-term heavy metal contamination. Environ. Pollut. 231: 908–917.
Liu, Z. and F. S. Zhang. 2009. Removal of lead from water using biochars prepared from hydrothermal liquefaction
of biomass. J. Hazard. Mater. 167: 933–939.
Luna, J. M., R. D. Rufino and L. A. Sarubbo. 2016. Biosurfactant from Candida sphaerica UCP0995 exhibiting heavy
metal remediation properties. Process Saf. Environ. Prot. 102: 558–566.
Mahdavian, K., S. M. Ghaderian and M. Torkzadeh-Mahani. 2017. Accumulation and phytoremediation of Pb, Zn,
and Ag by plants growing on Koshk lead–zinc mining area, Iran. J. Soils Sediments 17(5): 1310–1320.
Mani, D., C. Kumar, N. K. Patel and D. Sivakumar. 2015. Enhanced clean-up of lead-contaminated alluvial soil
through Chrysanthemum indicum L. Int. J. Environ. Sci. Technol. 12(4): 1211–1222.
Manzoor, M., A. Rathinasabapathi, L. M. De Oliveira, E. da Silva, F. Deng, C. Rensing, M. Arshad, G. Iram,
P. Xiang and L. Q. Ma. 2019. Metal tolerance of arsenic-resistant bacteria and their ability to promote plant
growth of Pteris vittata in Pb-contaminated soil. Sci. Total Environ. 660: 18–24. https://doi.org/10.1016/j.
scitotenv.2019.01.013.
204 Bioremediation for Sustainable Environmental Cleanup
Mariussen, E., I. V. Johnsen and A. E. Strømseng. 2018. Application of sorbents in different soil types from small
arms shooting ranges for immobilization of Lead (Pb), Copper (Cu), Zinc (Zn), and Antimony (Sb). J. Soil
Sediment. 18(4): 1558–68.
Meena, V., M. L. Dotaniya, J. K. Saha, H. Das and A. K. Patra. 2020. Impact of lead contamination on agroecosystem
and human health. pp. 67–82. In: D. K. Gupta, S. Chatterjee and C. Walther [eds.]. Lead in Plants and the
Environment. Switzerland: Springer Nature.
Mitra, A., S. Chatterjee, A. V. Voronina, C. Walther and D. K. Gupta. 2020. Lead toxicity in plants: a review.
pp. 99–116. In: D. K. Gupta, S. Chatterjee and C. Walther [eds.]. Lead in Plants and the Environment.
Switzerland: Springer Nature.
Mustapha, M. U. and N. Halimoon. 2015. Microorganisms and biosorption of heavy metals in the environment: a
review paper. J. Microb. Biochem. Technol. 7(5): 253–256.
Naik, M. M. and S. K. Dubey. 2017. Lead resistant bacteria: lead resistance mechanisms, their applications in lead
bioremediation and biomonitoring. Ecotoxicol. Environ. Saf. 98: 1–7.
Ojuederie, O. B. and O. O. Babalola. 2017. Microbial and plant-assisted bioremediation of heavy metal polluted
environments: a review. Int. J. Environ. Res. Public Health. 14(12): 1504.
Orroño, Daniela I., V. Schindler and R. S. Lavado. 2012. Heavy metal availability in Pelargonium hortorum
rhizosphere: interactions, uptake and plant accumulation. J. Plant Nutr. 35(9): 1374–86.
Panwar, N. R., J. K. Saha, T. Adhikari, S. Kundu, A. K. Iswas, A. Rathore, S. Ramana, S. Srivastava and A. Subba
Rao. 2010. Soil and water pollution in India: some case studies. IISS Technical Bulletin, Indian Institute of Soil
Science, Nabi Bagh, Bhopal, pp. 1–40.
Park, D., Y. S. Yun, J. H. Jo and J. M. Park. 2005. Mechanism of hexavalent chromium removal by dead fungal
biomass of Aspergillus niger. Water Res. 39(4): 533–540.
Patel, K. S., B. Ambade, S. Sharma, D. Sahu, N. K. Jaiswal, S. Gupta, R. K. Dewangan, S. Nava, F. Lucarelli,
B. Blazhev, R. Stefanova and J. Hoinkis. 2010. Lead environmental pollution in Central India. In: B. Ramov
[ed.]. New Trends in Technologies. IntechOpen.
Pipalde, J. S. and M. L. Dotaniya. 2018. Interactive effects of lead and nickel contamination on nickel mobility
dynamics in spinach. International Journal of Environmental Research 12(5): 553–560.
Pizzaia, D., M. L. Nogueira, M. Mondin, M. E. A. Carvalho, F. A. Piotto, M. F. Rosario and R. A. Azevedo. 2019.
Cadmium toxicity and its relationship with disturbances in the cytoskeleton, cell cycle and chromosome
stability. Ecotoxicol. 28(9): 1046–1055.
Pourrut, B., M. Shahid, F. Douay, C. Dumat and E. Pinelli. 2013. Molecular mechanisms involved in lead uptake,
toxicity and detoxification in higher plants. Heavy Metal Stress in Plants, 121–147.
Povedano-Priego, C., I. Martin-Sanchez, F. Jroundi, I. Sanchez-Castro and M. L. Merroun. 2017. Fungal
biomineralization of lead phosphates on the surface of lead metal. Miner. Eng. 106: 46–54.
Prabhakaran, P., M. A. Ashraf and W. S Aqma. 2016. Microbial stress response to heavy metals in the environment.
RSC Adv. 6(111): 109862–109877.
Prakash, D., P. Gabani, A. K. Chandel, Z. Ronen and O. V. Singh. 2013. Bioremediation: a genuine technology to
remediate radionuclides from the environment. Microbiol. Biotechnol. 6: 349–360. 10.1111/1751-7915.12059.
Punamiya, P., R. Datta, D. Sarkar, S. Barber, M. Patel and P. Das. 2010. Symbiotic role of Glomus Mosseae in
phytoextraction of lead in vetiver grass [Chrysopogon Zizanioides (L.)]. J. Hazard. Mater. 177(1–3): 465–74.
Rahman, Z. and V. P. Singh. 2020. Bioremediation of toxic heavy metals (THMs) contaminated sites: concepts,
applications and challenges. Environ. Sci. Pollut. Res. 27(22): 27563–27581. doi: 10.1007/s11356-020-08903-0.
Raj, D. and S. K. Maiti. 2020. Sources, bioaccumulation, health risks and remediation of potentially toxic metal(loid)s
(As, Cd, Cr, Pb and Hg): an epitomised review. Environ. Monitor. Assess. 192(2): 1–20.
Rhee, Y. J., S. Hillier and G. M. Gadd. 2012. Lead transformation to pyromorphite by fungi. Curr. Biol. 22: 237–241.
Rigoletto, M., P. Calza, E. Gaggero, M. Malandrino and D. Fabbri. 2020. Bioremediation methods for the recovery of
lead-contaminated soils: a review. App. Sci. 10(10): 3528. https://doi.org/10.3390/app10103528.
Romera, E., F. González, A. Ballester, M. L. Blázquez and J. A. Munoz. 2007. Comparative study of biosorption of
heavy metals using different types of algae. Bioresour. Technol. 98(17): 3344–3353.
Rucińska-Sobkowiak, R. 2016. Water relations in plants subjected to heavy metal stresses. Acta Physiologiae
Plantarum. 38(11): 1–13.
Rucińska-Sobkowiak, R., G. Nowaczyk, M. Krzesłowska, I. Rabęda and S. Jurga. 2013. Water status and water
diffusion transport in lupine roots exposed to lead. Environ. Exp. Botany. 87: 100–109.
Rufino, R. D., G. I. B. Rodrigues, G. M. Campos-Takaki, L. A. Sarubbo and S. R. M. Ferreira. 2011. Application
of a yeast biosurfactant in the removal of heavy metals and hydrophobic contaminant in a soil used as slurry
barrier. Applied and Environmental Soil Science, 2011.
Saha, J. K., N. Panwar and M. V. Singh. 2013. Risk assessment of heavy metals in soil of a susceptible agro-ecological
system amended with municipal solid waste compost. J. Indian Soc. Soil Sci. 61: 15–22.
Lead Induced Toxicity, Detoxification and Bioremediation 205
Sahmoune, M. N. 2018. Performance of Streptomyces rimosus biomass in biosorption of heavy metals from aqueous
solutions. Microchem. J. 141: 87–95.
Salama, A. K., K. A. Osman and N. A. R. Gouda. 2016. Remediation of lead and cadmium-contaminated soils. Int. J.
Phytoremediation. 18(4): 364–367.
Saleh, H. M., R. F. Aglan and H. H. Mahmoud. 2019. Ludwigia stolonifera for remediation of toxic metals from
simulated wastewater. Chem. Ecol. 35(2): 164–178.
Sammut, M. L., Y. Noack, J. Rose, J. L. Hazemann, O. Proux, M. Depoux, A. Ziebel and E. Fiani. 2010. Speciation
of Cd and Pb in dust emitted from sinter plant. Chemosphere. 78(4): 445–50.
Santos, C. L., B. Pourrut and J. M. P. Oliveira. 2015. The use of comet assay in plant toxicology: recent
advances. Frontiers in Genetics. 6: 216.
Shahid, M., C. Dumat, S. Khalid, E. Schreck, T. Xiong and N. K. Niazi. 2017. Foliar heavy metal uptake, toxicity
and detoxification in plants: a comparison of foliar and root metal uptake. Journal of Hazardous Materials.
325: 36–58.
Shahid, M., E. Pinelli, B. Pourrut, J. Silvestre and C. Dumat. 2011. Lead-induced genotoxicity to vicia faba l. roots in
relation with metal cell uptake and initial speciation. Ecotoxicol. Environ. Saf. 74(1): 78–84.
Shi, L., H. Dong, G. Reguera, H. Beyenal, A. Lu, J. Liu, H. Q. Yu and J. K. Fredrickson. 2016. Extracellular electron
transfer mechanisms between microorganisms and minerals. Nat. Rev. Microbiol. 14: 651–662. https://doi.
org/10.1038/nrmicro.2016.93.
Sidhu, G. P. S., A. S. Bali, H. P. Singh, D. R. Batish and R. K. Kohli. 2018. Phytoremediation of lead by a wild, non-
edible Pb accumulator Coronopus didymus (L.) Brassicaceae. Int. J. Phytoremediation. 20(5): 483–489.
Singh, A. and S. M. Prasad. 2015. Remediation of heavy metal contaminated ecosystem: an overview on technology
advancement. Int. J. Environ. Sci. Technol. 12 (1): 353–366. doi:10.1007/s13762-014-0542-y.
Solgi, E., R. Yazdanyar and M. Taghizadeh. 2020. Assessment of phytoremediation potential of Alyssum Maritimum
in remediation of lead-contaminated soils. J. School Pub. Health Inst. Public Health Res. 17(4): 363–372.
Srivastava, S., and I. S. Thakur. 2006. Isolation and process parameter optimization of Aspergillus sp. for removal of
chromium from tannery effluent. Bioresource Technology. 97(10): 1167–1173.
Su, C. 2014. A review on heavy metal contamination in the soil worldwide: situation, impact and remediation
techniques. Environ. Skeptics and Critics. 3: 24–38.
Sylvain, B., M. H. Mikael, M. Florie, J. Emmanuel, S. Marilyne, B. Sylvain and M. Domenico. 2016. Phytostabilization
of As, Sb and Pb by two willow species (S. viminalis and S. purpurea) on former mine technosols. Catena.
136: 44–52.
Tayang A. and L. S. Songachan. 2021. Microbial bioremediation of heavy metals. Current. Sci. 120(6): 1013–1025.
Thakare, M., H. Sarma, S. Datar, A. Roy, P. Pawar, K. Gupta, S. Pandit and R. Prasad. 2021. Understanding the
holistic approach to plant-microbe remediation technologies for removing heavy metals and radionuclides
from soil. Cur. Res. Biotech. 3: 84–98.
Tiwari, S., S. N. Singh and S. K. Garg. 2013. Microbially enhanced phytoextraction of heavy-metal fly-ash amended
soil. Communications in Soil Science and Plant Analysis. 44(21): 3161–3176.
Ullah, A., S. Heng, M. F. H. Munis, S. Fahad and X. Yang. 2015. Phytoremediation of heavy metals assisted by plant
growth promoting (PGP) bacteria: a review. Environ. Exp. Bot. 117: 28–40.
Vacar, C. L., E. Covaci, S. Chakraborty, B. Li, D. C. Weindorf, T. Frentiu, M. Parvu and D. Podar. 2021. Heavy metal-
resistant filamentous fungi as potential mercury bioremediators. J. Fungi. 7: 386.
Vega, F. A., M. L. Andrade and E. F. Covelo. 2010. Influence of soil properties on the sorption and retention of
cadmium, copper and lead, separately and together, by 20 soil horizons: comparison of linear regression and
tree regression analyses. J. Hazard. Mater. 174 (1–3): 522–33.
Waranusantigul, P., H. Lee, M. Kruatrachue, P. Pokethitiyook and C. Auesukaree. 2011. Isolation and characterization
of lead-tolerant Ochrobactrum intermedium and its role in enhancing lead accumulation by Eucalyptus
camaldulensis. Chemosphere. 85(4): 584–590.
Wu, Y., Q. Liang and Q. Tang. 2011. Effect of Pb on growth, accumulation and quality component of tea plant.
Procedia Eng. 18: 214–19.
Wuana, R. A. and F. E. Okieimen. 2011. Heavy metals in contaminated soils: a review of sources, Chemistry, Risks
and Best Available Strategies for Remediation. ISRN Ecol. 2011: 1–20.
Xie, Y., J. Fan, W. Zhu, E. Amombo, Y. Lou, L. Chen and J. Fu. 2016. Effect of heavy metals pollution on soil
microbial diversity and bermudagrass genetic variation. Frontiers in Plant Science, 7: 755.
Yalçın, S., S. Sezer and R. Apak. 2012. Characterization and lead (II), cadmium (II), nickel (II) biosorption of dried
marine brown macro algae Cystoseira barbata. Environmental Science and Pollution Research. 19(8): 3118–
3125.
Yang, J., J. Yang and J. Huang. 2017. Role of co-planting and chitosan in phytoextraction of As and heavy metals by
Pteris vittata and castor bean–a field case. Ecol. Eng. 109: 35–40.
206 Bioremediation for Sustainable Environmental Cleanup
Yang, T., M. L. Chen and J. H. Wang. 2015. Genetic and chemical modification of cells for selective separation and
analysis of heavy metals of biological or environmental significance. TrAC Trends in Analytical Chemistry.
66: 90–102.
Yang, Y., L. Zhang, X. Huang, Y. Zhou, Q. Quan, Y. Li and X. Zhu. 2020. Response of photosynthesis to different
concentrations of heavy metals in Davidia involucrata. PloS one. 15(3): 0228563.
Yin, K., M. Lv, Q. Wang, Y. Wu, C. Liao, W. Zhang and L. Chen. 2016. Simultaneous bioremediation and biodetection
of mercury ion through surface display of carboxylesterase E2 from Pseudomonas aeruginosa PA1. Water
Res. 103: 383–390.
Yin, K., Q. Wang, M. Lv and L. Chen. 2019. Microorganism remediation strategies towards heavy metals. Chem.
Eng. J. 360: 1553–1563.
Zaier, H., T. Ghnaya, R. Ghabriche, W. Chmingui, A. Lakhdar, S. Lutts and C. Abdelly. 2014. EDTA-enhanced
phytoremediation of lead-contaminated soil by the halophyte Sesuvium portulacastrum. Environ. Sci. Pollut.
Res. 21(12): 7607–7615.
Zhao, L., T. Li, X. Zhang, G. Chen, Z. Zheng and H. Yu. 2016. Pb Uptake and phytostabilization potential of the
mining ecotype of Athyrium wardii (Hook.) grown in Pb‐Contaminated soil. Clean–Soil Air Water. 44(9):
1184–1190.
Zulfiqar, U., M. Farooq, S. Hussain, M. Maqsood, M. Hussain, M. Ishfaq, M. Ahmad and M. Z. Anjum. 2019. Lead
toxicity in plants: impacts and remediation. J. Environ. Manag. 250: 109557.
Chapter 12
Microalgal Bioremediation of
Heavy Metals:
An Integrated Low-cost Sustainable
Approach
Anubha Kaushik,1 Sharma Mona,2,3,* Randhir Bharti1
and Sujata 3
12.1 Introduction
Microalgae consisting of green algae and cyanobacteria have a remarkable range of applications,
including bioremediation of pollutants and production of biofuel, bioactive compounds and
biofertilizers. Microalgae have a short life cycle, and can be easily cultured in the laboratory.
Microalgae may be grown either in the conventional suspended form or in the attached form. The
suspended form has relatively lower efficiency for bioremediation, whereas, in the attached form,
they tend to form a biofilm that is more effectual in wastewater treatment and bioremediation
(Hasan et al. 2021). With a broad range of use for various heavy metals in different sectors, their
concentrations in wastewaters are increasing at an alarming rate in the wastewaters and solid wastes,
which ultimately enter the aquatic and terrestrial ecosystems, pollute the environment and health
is also affected (Briffa et al. 2020). There is a need to eliminate toxic metals from the environment
to safe levels for which bioremediation has emerged as an environmentally sound and low-cost
approach. The part of microalgae in the bioremediation of various toxic metals, various processes
and mechanisms, and the possibilities of adopting a biorefinery approach for wastewater treatment
with microalgae have been described in the chapter to make it an economically and environmentally
sustainable method.
1
University School of Environment Management, GGS Indraprastha University, New Delhi-110078, India.
2
Department of Environmental Studies, Central University of Haryana, Mahendergarh-123031, Haryana, India.
3
Department of Environmental Science & Engineering, Guru Jambheshwar University of Science & Technology,
Hisar-125001, Haryana, India.
* Corresponding author: [email protected], [email protected]
208 Bioremediation for Sustainable Environmental Cleanup
1.Ultrafiltration 1.Bioaccumulation
1. Neutralization
2.Coagulation 2. Biosorption
2. Solvent extraction
3.Flocculation 3. Microbial
3. Chemical
reduction of
precipitation
4.Membrane oxidation
filtration 4. Electrochemical
4. Metabolic
treatment
5.Ion exchange precipitation
5. Metal-
Phytochelatin
By using physical methods, almost all the pollutants can be removed, but these methods have
some limitations. Physical methods based on the distribution of the practical size of pollutants need
further processing and have a comparatively high cost of application. Although, chemical methods
of metal remediation are highly effective in these methods formation of byproducts increases further
downstream processing steps (Mona et al. 2008). The biological methods for metal remediation are
less costly and do not create any secondary pollution (Selvi et al. 2019).
Some metals like Cu, Ni, Mn, Fe are essential for the growth and conditioning of microorganisms
as they are required for the cells (Toranzo et al. 2020). While metals like Pb, Cd, Hg and As are
harmful at very low concentrations, essential metals are also toxic for microbial activity beyond
a threshold level (Hall 2002). Essential metals are displaced by non-essential metals because of
their high ionic force from their area of bindings and make complex bindings with working sites of
microbial cell walls.
of researchers across the world, leading to a huge amount of literature on the subject. Table 12.3
shows some reports on metal removal by various microalgae species.
Bioaccumulation is a procedure wherein heavy metal is controlled metabolically, producing
energy and altering heavy metal concentration (Arunakumara and Zhang 2008). Bioaccumulation
of metals by living cells depends on both the intra and extracellular processes, and passive gaining
is restricted (Fomina and Gadd 2014). Cladophora herpestica, a green alga that grows abundantly
in the Maruit Lake surface, has shown accumulation of residual nutrients from the atmospheric
and aquatic environment in introduction to heavy metal ions (Dahlia and Hassan 2017). Some
microalgae have also been reported to heavy metals bioaccumulation. Bioaccumulation of heavy
metals (Zn, Fe, Cu, Cd,Als) was shown by Chlorella vulgaris, Phacus curvicauda, Euglena acus
and Oscillatoria bornettia (Abrihire and Kadiri 2011). Amongst these species, Oscillatoria had a
high concentration of metal factor for Zn (0.306), Fe (0.302), Cu (0.091), Cd (0.276), while Phacus
and Euglena had relatively higher concentration factor for Al (0.439).
The temperature has the highest effect on metal remediation, but it has less impact than the
impact of pH (Lau et al. 1999). Therefore, it is important to find the optimum temperature for
each microalgal system before designing the biosorption operation for different metals.
(c) Organic matter: Organic matter concentration influences the biosorption method, and is
dependent on the type of metal and microalgal species used in bioremediation. For example,
biosorption of copper decreases from 3.2–2.3 mg L–1 and arsenic from 2.2–0.0 mg L–1 in
Chlorella, whereas in the case of Scendesmus almeriensis, removal declines from 2.1 to
1.6 mg L–1 for copper and 2.3 to 1.7 mg L–1 for arsenic in the existence of organic matter
(Saavedra et al. 2019). Since most wastewaters have some organic load, and organic matter
tends to chelate the metals-forming complexes, therefore it is paramount to comprehend the
effect of organic matter on the overall metal bioremediation protocol.
(d) Carbon dioxide: Microalgae need carbon dioxide for their photosynthesis as raw material, and
carbon dioxide is emitted by several industries. Microalgae can be grown by using this waste
gas and converting the same into useful biomass, which can then be used as a biosorbent to
clean metal contaminated waste streams. Oedogonium sp. shows rapid uptake of a few heavy
metals (particularly Cd, Al, Ni and Zn), and its biomass productivity gets increased when CO2
concentration is higher than normal conditions and escalates with productivity (Roberts et al.
2013).
(e) Nutrients: Algal growth is largely affected by the availability of nutrients. There are some
dominant species likes Chlamydomonas, Spirogyra, Euglena and Dinoflagellates that use
carbon and other nutrients from the wastewater for their maturation and photosynthetic activity,
increasing the efficiency of metal remediation microorganisms (Sayara et al. 2021).
Many other factors like initial metal concentration, types of metals such as tertiary and
quaternary multi-metal systems and contact time are also important factors for metal removal (Kiran
and Kaushik 2012).
transform infrared, spectroscopy (IR or FTIR), X-ray Absorption Spectroscopy (XAS) and Nuclear
Magnetic Resonance (NMR) to gain insight into the process and mechanisms of biosorption process
(Fomina and Gadd 2014, Michalak et al. 2013, Kiran et al. 2016).” It can be confirmed that surface
adsorption of the metals is found in the case of Cr (VI) bound at specific binding sites on the algal
biosorbent based on SEM and FTIR spectral analysis (Kiran et al. 2008).
To understand the procedure of the biosorption and optimize the processes, modeling and
simulation of the experimental data are generally done for which several models have been
developed (Volesky 2003a). The two regularly used models are Langmuir and Freundlich models, in
which the adsorption mechanism is demonstrated as a batch equilibrium isotherm curve to contrast
pollutant uptake proportion of different bio-sorbent and affinities for the metals (Mona et al. 2011a).
These equilibrium sorption models provide some normal information about a given process. As
the biosorbent adsorbs the metal to the equilibrium point, the value of metal uptake (qe) by the
bio-sorbent is plotted against the equilibrium (final) metal concentration (C). The Langmuir isotherm
assume a finite number of equal adsorption locations and the absence of lateral interrelation between
adsorbed species. The most regularly multilayer adsorption model is the Brunauer–Emmett–Teller
(BET) isotherm, which presume that the Langmuir equation is applied to every layer (Vijayaraghavan
and Yun 2008). Equilibrium data is right to Langmuir isotherm, and a linear plot is received from
this isotherm (Figure 12.2); and, the value of the Langmuir constant is calculated:
qe = QobCe + bCe
where;
qe = adsorption of metal (mg g–1); Ce = residual metal (mg L–1)
Qo (mg g–1) and b (L mg–1) are Langmuir constants exhibit the adsorption range and adsorption
energy (Mona et al. 2015a).
Freundlich isotherm considers a heterogeneous base of the adsorbent, and the equation is
“where;
qe = metal adsorbed (mg g–1); Ce = residual metal ion concentration (mg L–1); n = Freundlich
exponent; Kf = Freundlich constant indicating adsorbent capacity (mg g–1 dry weight).”
And a linear plot of Log qe versus Log Ce explains the applicability of this isotherm (Figure
12.2) for the biosorbent (Gadd 2009).
There is another model, Brunauer Emmer and Teller (BET); in this model the 1st layer of
sorbent gets absorbed on the upper layer with the energy approximate to the heat of adsorption for
single layer sorption, and the next layer has same energy.
Ce 1 Ce
qe = +
qmKL qm
“where;
Ce = the equilibrium concentration of the adsorbate and qe is the adsorption capacity adsorbed at
equilibrium, qm is maximum adsorption capacity
KL = the Langmuir adsorption constant”
Response Surface Methodology (RSM) is one of the most widely used approaches adopted for
optimization of the parameters to get the maximum metal removal response from the biosorption
process (Mona et al. 2011a). Box-Behnken Design (BBD) and Central Composite Design (CCD) are
two models of Response Surface Methodology (RSM). Box-Behnken model (BBM) of RSM is used
214 Bioremediation for Sustainable Environmental Cleanup
Figure 12.2. Equations and graphical representation of Freundlich, Langmuir and Brunuaer–Emmett–Teller (BET)
Figure.(adopted
isotherms 12.2. Equations and
from Gadd 2009, graphical
Park et al. 2011).representation of Freundlich,
Langm
and Brunuaer–Emmett–Teller (BET) isotherms (adopted from Gadd, 2009;
in the study with maximumetand al. minimum
2011). levels for each parameter (initial metal concentration,
pH, temperature). The CCD methodology is used for the prediction of the impact of different
parameters on the metal remediation process (Podstawczyk et al. 2015). To regulate the adsorption
equilibrium and kinetic behavior of chromium in an aqueous solution, batch mode experiments
were performed for Lyngbya, and maximum metal adsorption capacity was determined. The impact
of other parameters like initial metal ion concentration (10–100 mg/L), pH (2–6) and temperature
(25–45°C) on chromium adsorption were also noted , using Box–Behnken design (BBD) model of
RSM and the optimized conditions (50–60 mg/L initial Cr concentration, 2–3 pH, 45°C temperature)
were computed, when 82% of the metal was extracted from the solution by the alginate immobilized
alga (Kiran et al. 2007b).
The biosorption range of algal cells directly depends on the presence of different functional
groups viz. OH–, PO3O2, NH2, COOH, SH, etc., on the cell of algae. These groups provide a negative
charge to the cell (Kaplan 2013), which has a strong binding capacity for the positively charged
heavy metals present in the water. These functional groups form a complex or get linked with cell
wall components (teichoic acids, peptidoglycan, polysaccharides and proteins), directly or indirectly
provide metal-binding sites (Volesky and Holan 1995). Biosorption of Cd in Chlamydomonas
reinhardtti occurs in a unique manner, where Cd2+ ions form a complex with carboxylic groups
of the algal cells (Adhiya et al. 2002). Covalent bonding and electrostatic attraction are directly
responsible for the adsorption of Ni and Zn on the Chaetophora elegans (Andrade et al. 2005).
Bioremediation of aluminum (Al) follows a different mechanism of sorption on algal cells as it
Microalgal Bioremediation of Heavy Metals 215
binds to the algal cell in a polynuclear form; after binding of aluminum ions, these ions prevent the
other metal ions from binding on the binding surface of the biomass (Bottero et al. 1980).
microalga has been demonstrated with high biosorption capacities 14.3mg Cr g–1 EPS and
17.9 mg Co g–1 EPS (Mona and Kaushik 2015b).
BV SORBEX, a company owned by the pioneer researcher in the field of biosorption, McGill
University and Bohumil Volesky from Canada, has applied the laboratory process of biosorption to
an industrial level (Volesky 2003b).
However, there are some limitations in the application of the biosorption method on an industrial
scale. This limitation is due to less durability and low mechanical resistance of the biomass, which
create problems in the process and other factors, including regeneration of the sorbent and its
subsequent declination. Thus the success of the sorption process depends on the recovery and reuse
of the sorbent. It is further required that the biosorption process be integrated into other microalgae
based applications, including energy recovery in the form of biofuels like biohydrogen or biodiesel
and value-added biomaterial production to make the overall procedure more economically feasible
and sustainable. Next new directions for integrating various microalgae-based technologies have
been described with a view to make it a competitive metal bioremediation technology with additional
applications to raise its sustainability. Various new opportunities for such combined technologies
and the biorefinery approach for microalgal wastewater treatment and bioremediation are discussed
next.
adsorption capacity (15–18 mg/g) for Cr (VI) and Cd metals, and the efficiency is higher when the
spent biomass is obtained at a relatively earlier stage (Mona et al. 2013). Such studies suggest that
careful planning can lead to the highly successful integration of the algae-based bioremediation
cum-bioenergy production.
These studies explore the application of waste algal biomass from biofuel producing systems for
metal bioremediation and use microalgal biomass produced in wastewater for biofuel production.
References
Abrihire, O. and M. O. Kadiri. 2011. Bioaccumulation of heavy metals using microalgae. Asian J. Microbiol.
Biotechol. Environ. Sci. 13: 91–94.
Adhiya, J., X. Cai, R. T. Sayre and S. J. Traina. 2002. Binding of aqueous cadmium by the lyophilized biomass of
Chlamydomonas reinhardtii. Colloid. Surf. A: Physicochem Eng. Asp. 210: 1–11.
Ahluwalia, S. S. and D. Goyal. 2007. Microbial and plant derived biomass for removal of heavy metals from
wastewater. Bioresour. Technol. 2243–2257.
Aksu, Z. 2002. Determination of the equilibrium, kinetic and thermodynamic parameters of the batch biosorption of
nickel (II) ions onto Chlorella vulgaris. Process Biochem. 38(1): 89–99.
Andrade, A. D., M. C. E. Rollemberg and J. A. Nóbrega. 2005. Proton and metal binding capacity of the green
freshwater alga Chaetophora elegans. Process Biochem. 40: 1931–1936.
Arunakumara, K. K. I. U. and X. Zhang. 2008. Heavy metal bioaccumulation and toxicity with special reference to
microalgae. J. Ocean Uni. China. 7: 60–64.
Bhatia, D., N. R. Sharma, J. Singh and R. S. Kanwar. 2017. Biological methods for textile dye removal from
wastewater: a review. Crit. Rev. Environ. Sci. Technol. 47: 1836–1876.
Bottero, J. Y., J. M. Cases, F. Fiessenger and J. Poirier. 1980. Studies of hydrolysed Aluminium Chloride solutions.
Nature of Aluminium species and composition of aqueous solutions. J. Phys. Chem. 84: 2933–2939.
Briffa, J., E. Sinagra and R. Blundell. 2020. Heavy metal pollution in the environment and their toxicological effects
on humans. Heliyon. 6: e04691.
Brinza, L., C. A. Nygård, M. J. Dring, M. Gavrilescu and L. G. Benning. 2009. Cadmium tolerance and adsorption
by the marine brown alga Fucus vesiculosus from the Irish Sea and the Bothnian Sea. Bioresour. Technol.
100(5): 1727–1733.
Cameron, H., M. T. Mata and C. Riquelme. 2018. The effect of heavy metals on the viability of Tetraselmis marina
AC16-MESO and an evaluation of the potential use of this microalga in bioremediation. Peer. 6: e5295.
Craggs, R. J., S. Heubeck, T. J. Lundquist and J. R. Benemann. 2011. Algal biofuels from wastewater treatment high
rate algal ponds. Water Sci. Technol. 63: 660–665.
Dahlia, M. and I. Hassan. 2017. Heavy metals bioaccumulation by the green alga Cladophera herpestica in Lake
Mariut, Alexandria, Egypt. J. Pollut. 1: 2.
Demayo, A., M. C. Taylor, K. W. Taylor, P. V. Hodson and P. B. Hammond. 1982. Toxic effects of lead and lead
compounds on human health, aquatic life, wildlife plants, and livestock. Crit. Rev. Environ. Sci. Technol.
12(4): 257–305.
Duncan, J. R., D. Brady and B. Wilhelmi. 1997. Immobilization of yeast and algal cells for bioremediation of heavy
metals. In Biorem. Proto. 91–97.
Eccles, H. 1999. Treatment of metal-contaminated wastes: why select a biological process. Trends Biotechnol.
17: 462–465.
El-Sheekh, M. M., A.A. Farghl, H. R. Galal and H. S. Bayoumi. 2016. Bioremediation of different types of polluted
water using microalgae. Rendiconti Lincei, 27: 401–410.
Fatima, G., A. M. Raza, N. Hadi, N. Nigam and A. A. Mahdi. 2019. Cadmium in human diseases: it’s more than just
a mere metal. Indian J. Clin. Biochem. 34: 371–378.
Fomina, M. and G. M. Gadd. 2014. Biosorption: current perspectives on concept, definition and application.
Bioresour. Technol. 160: 3–14.
Gadd, G. M. 2009. Biosorption: critical review of scientific rationale, environmental importance and significance
for pollution treatment. Journal of Chemical Technology & Biotechnology: International Research in Process.
Environmental & Clean Technology. 84(1): 13–28.
García-García, J. D., K. A. Peña-Sanabria, R. Sánchez-Thomas and R. Moreno-Sánchez. 2018. Nickel accumulation
by the green algae-like Euglena gracilis. J. Hazard. Mater. 343: 10–18.
García-Lestón, J., J. Méndez, E. Pásaro and B. Laffon. 2010. Genotoxic effects of lead: an updated review. Environ.
Int. 36(6): 623–636.
Gaur, J. P. and L. C. Rai. 2001. Heavy metal tolerance in algae. Algal adaptation to environmental stresses. Springer,
Berlin, Heidelberg, 363–388.
Genchi, G., A. Carocci, G. Lauria, M. S. Sinicropi and A. Catalano. 2020. Nickel: human health and environmental
toxicology. Int. J. Environ. Res. Public Health, 17: 679.
Gonçalves A. L., J. C. Pires and M. Simões. 2017. A review on the use of microalgal consortia for wastewater
treatment. Algal Res. 24: 403–415.
220 Bioremediation for Sustainable Environmental Cleanup
Gong R., Y. Ding, H. Liu, Q. Chen and Z. Liu. 2005. Lead biosorption and desorption by intact and pretreated
Spirulina maxima biomass. Chemosphere. 58: 125–130.
Hall, J. Á. 2002. Cellular mechanisms for heavy metal detoxification and tolerance. J. Exp. Bot. 53: 1–11.
Hamdy, A. A. 2000. Biosorption of heavy metals by marine algae. Current Microbiol. 41: 232–238.
Hasan, H. A., S. N. Hatika, A. Bakar and M. S. Takriff. 2021. Microalgae biofilms for the treatment of
wastewater. Microalgae. Academic Press, 2021. 381–407.
Igiri, B. E., S. I. Okoduwa, G. O. Idoko, E. P. Akabuogu, A. O. Adeyi and I. K. Ejiogu. 2018. Toxicity and bioremediation
of heavy metals contaminated ecosystem from tannery wastewater: a review. Journal of Toxicology.
Jankowska, E., K. S. Ashish and O. P. Piotr. 2017. Biogas from microalgae: review on microalgae’s cultivation,
harvesting and pretreatment for anaerobic digestion. Renew. Sustain. Energy Rev. 75: 692–709.
Kalra, R., S. Gaur and M. Goel. 2020. Microalgae bioremediation: a perspective towards wastewater treatment along
with industrial carotenoids production. J. Water Process Eng. 101794.
Kamra, A., A. Kaushik, K. Bala and N. Rani. 2007. Biosorption of Cr(VI) by immobilized biomass of two indigenous
strains of cyanobacteria isolated from metal contaminated soil. J. Hazard. Mater. 148: 383–386.
Kaplan, D. 2013. Absorption and adsorption of heavy metals by microalgae. Handbook of microalgal culture: applied
Phycology and Biotechnology. 2: 602–611.
Kaushik, A., S. Jain, J. Dawra and M. S. Bishnoi. 2000. Heavy metal pollution of river Ghaggar in Haryana. Indian
J. Environ. Toxicol. 10: 63–66.
Kaushik, A., S. Jain, J. Dawra and C. P. Kaushik. 2001. Heavy metal pollution of river Yamuna in the industrially
developing state of Haryana. J. Environ. Biol. 43: 164–168.
Kaushik, A., S. Mona and C. P. Kaushik. 2011. Integrating photobiological hydrogen production with dye-metal
bioremoval from simulated textile wastewater. Bioresour. Technol. 102: 9957–9964.
Kiran, B., A. Kaushik and C. P. Kaushik. 2007a. Biosorption of Cr (VI) by native isolate of L. putealis (HH-15) in the
presence of salts. J. Hazard. Mater. 141: 662–667.
Kiran, B., A. Kaushik and C. P. Kaushik. 2007b. Response surface methodological approach for optimizing removal
of Cr (VI) from aq. sol. using immobilized cyanobacterium. Chem. Eng. J. 126: 147–153.
Kiran, B. and A. Kaushik. 2008. Cyanobacterial biosorption of Cr (VI): application of two parameter and Bohart
Adams models for batch and column studies. Chem. Eng. J. 144: 391–399.
Kiran, B., A. Kaushik and C. P. Kaushik. 2008. Metal–salt co-tolerance and metal removal by indigenous cyanobacterial
strains. Process Biochemistry. 43.6: 598–604.
Kiran, B. and A. Kaushik. 2012. Equilibrium sorption study of Cr (VI) from multi-metal systems in aq. sol. by
Lyngbya putealis. Ecol. Eng. 38: 93–96.
Kiran, B., N. Rani and A. Kaushik. 2016. FTIR spectroscopy and scanning electron microscopic analysis of pretreated
biosorbent to observe the effect on Cr (VI) remediation. Int. J. Phytoremediat. 18: 1067–1074.
Kosek, K., Z. Polkowska, B. Żyszka and J. Lipok. 2016. Phytoplankton communities of polar regions diversity
depending on enviro. conditions and chemical anthropopressure. J. Environ. Manag. 171: 243–259.
Kratochvil, D. and B. Volesky. 1998. Biosorption of Cu from Ferruginous Wastewater by Algal Bio-mass. Water Res.
32: 2760–2768.
Kumar, D., L. K. Pandey and J. P. Gaur. 2016. Metal sorption by algal biomass: from batch to continuous system.
Algal Res. 18: 95–109.
Kumar, K. S., H. U. Dahms, E. J. Won, J. S. Lee and K. H. Shin. 2015. Microalgae–a promising tool for heavy metal
remediation. Ecotoxicol. Enviro Saf. 113: 329–352.
Kumar, V., A. Sharma and A. Cerda. 2020. Heavy Metals in the Environment: Impact, Assessment, and Remediation
Elsevier.
Lau, P. S., H. Y. Lee, C. C. K. Tsang, N. F. Y. Tam and Y. S. Wong. 1999. Effect of metal interference, pH and
temperature on Cu and Ni biosorption by C. Vulgaris and C. Miniata. Environ. Technol. 20: 953–96.
Li, Y., Y. Chen and J. Wu. 2019. Enhancement of methane production in anaerobic digestion process: A review.
Applied Energy 240: 120–137.
Markou, G. and E. Nerantzis. 2013. Microalgae for high-value compounds and biofuels production: a review with
focus on cultivation under stress conditions. Biotechnol. Adv. 31: 1532–1542.
Michalak, I., K. Chojnacka and A. Witek-Krowiak. 2013. State of the art for the biosorption process—a review. Appl.
Biochem. Biotechnol. 170: 1389–1416.
Mona, S., A. Kaushik and C. P. Kaushik. 2011a. Biosorption of Chromium (VI) by spent cyanobacterial biomass from
a hydrogen fermentor using box-behnken model. Int. Biodeterior. Biodegrad. 65: 656–663.
Mona, S., A. Kaushik and C. P. Kaushik. 2011b. Sequestration of Co (II) from aqueous solution using immobilized
biomass of Nostoc linckia waste from a hydrogen bioreactor. Desalin. 276: 408–415.
Mona, S., A. Kaushik and C. P. Kaushik. 2011c. Hydrogen production and metal-dye bioremoval by a N. linckia strain
isolated from textile mill oxidation pond. Bioresour. Technol. 102: 3200–3205.
Microalgal Bioremediation of Heavy Metals 221
Mona, S., A. Kaushik and C. P. Kaushik. 2013. Prolonged hydrogen production by Nostoc in photobioreactor and
multi-stage use of the biological waste for column biosorption of some dyes and metals. Biomass Bioenergy
54: 27–35.
Mona, S. and A. Kaushik. 2015a. Screening metal-dye-tolerant photoautotrophic microbes from textile wastewaters
for biohydrogen production J. Appl. Phycol. 27: 1185–1194.
Mona, S. and A. Kaushik. 2015b. Cr and Cb sequestration using exopolysaccharides produced by freshwater
cyanobacterium Nostoc linckia. Ecol. Eng. 82: 121–125.
Monteiro, C., P. Castro and F. Malcata. 2012. Metal uptake by microalgae: underlying mechanisms and practical
applications. Biotechnol. Progress. 28: 299–311.
Naimabadi, A., A. Gholami and A. M. Ramezani. 2020. Determination of heavy metals and health risk assessment
in indoor dust from different functional areas in Neyshabur, Iran. Indoor Built Environ. 1420326X20963378.
Nasirpour, N., S. M. Zamir and S. A. Shojaosadati. 2017. Immobilization techniques for microbial bioremediation of
toxic metals. In Handbook of Metal-microbe Interactions and Bioremed. 357–370.
Park, J. B, K. R. J. Craggs and A. N. Shilton. 2011. Wastewater treatment high rate algal ponds for biofuel production.
Bioresour. Technol. 102: 35–42.
Picardo, M. C., J. L. de Medeiros, F. A. Ofélia de Queiroz and R. M. Chaloub. 2013. Effects of CO2 enrichment and
nutrients supply intermittency on batch cultures of Isochrysis galbana. Bioresour. Technol. 143: 242–250.
Podstawczyk, D., A. Witek-Krowiak, A. Dawiec and A. Bhatnagar. 2015. Biosorption of copper (II) ions by flax meal:
empirical modeling and process optimization by response surface methodology (RSM) and artificial neural
network (ANN) simulation. Ecol. Eng. 83: 364–379.
Potnis, A. A., P. S. Raghavan and H. Rajaram. 2021. Overview on cyanobacterial exopolysaccharides and biofilms:
role in bioremediation. Reviews in Environmental Science and Bio/Technology. 20.3: 781–794.
Preeti, R., A. Kaushik and C. P. Kaushik. 2015. Potential of cyanobacterial consortium of hot springs for high value
pigment production and dye decolorization. Inter. J. Adv. Sci. Technol. 53: 11–317.
Priyadarshini, E., S. S. Priyadarshini and N. Pradhan. 2019. Heavy metal resistance in algae and its application for
metal nanoparticle synthesis. Appl. Microbiol. Biotechnol. 103: 3297–3316.
Reed, R. H. and G. M. Gadd. 1989. Metal tolerance in eukaryotic and prokaryotic algae. Heavy metal tolerance in
plants: Evolutionary Aspects. 105–118.
Rehman, K., F. Fatima, I. Waheed and M. S. H. Akash. 2018. Prevalence of exposure of heavy metals and their impact
on health consequences. J. Cell. Biochem. 119: 157–184.
Roberts, D. A., R. de Nys and N. A. Paul. 2013. The effect of CO2 on algal growth in industrial waste water for
bioenergy and bioremediation applications. PloS one. 8: e81631.
Saavedra, R., R. Muñoz, M. E. Taboada and S. Bolado. 2019. Influence of organic matter and CO2 supply on
bioremediation of heavy metals by C. vulgaris and S. almeriensis in a multi metallic matrix. Ecotoxicol.
Environ. Saf. 182: 109393.
Sayara, T., S. Khayat, J. Saleh and P. Van Der Steen. 2021. Evaluation of the effect of reaction time on nutrients
removal from Sec. effluent of wastewater: Field demonstrations using algal-bacterial photobioreactors. Saudi
J. Biol. Sci. 28: 504–511.
Schenk, P. M., S. R. Thomas-Hall, E. Stephens, U. C. Marx, J. H. Mussgnug, C. Posten, O. Kruse and B. Hankamer.
2008. Second generation biofuels: high-efficiency microalgae for biodiesel production. Bioenergy Res.
1: 20–43.
Selvi, A., A. Rajasekar, J. Theerthagiri, A. Ananthaselvam, K. Sathishkumar, J. Madhavan and P. K. Rahman.
2019. Integrated remediation processes toward heavy metal removal/recovery from various environments-a
review. Frontiers Environ. Sci. 7: 66.
Sharma, M., A. Kaushik, K. Bala and A. Kamra. 2008. Sequestration of chromium by exopolysaccharides of Nostoc
and Gloeocapsa from dilute aqueous solutions. Journal of Hazardous Materials 157(2-3): 315–318.
Sharma, M., N. Rani, A. Kamra, A. Kaushik and K. Bala. 2009. Growth, exopolymer production and metal bioremoval
by Nostoc punctiforme in Na and Cr (VI) spiked medium. J. Environ. Res. Dev. 4(2).
Sharma, S., S. Rana, A. Thakkar, A. Baldi, R. S. R. Murthy and R. K. Sharma. 2016. Physical, chemical and
phytoremediation technique for removal of heavy metals. J. Heavy Met. Toxic. Dis. 1: 1–15.
Singh, A., D. Kumar and J. P. Gaur. 2012. Continuous metal removal from solution and industrial effluents
using Spirogyra biomass-packed column reactor. Water Res. 46(3): 779–788.
Sweetly, K., J. Sangeetha and B. Suganthi. 2014. Biosorption of heavy metal lead from aq. Sol by non- living biomass
of Sargassum myriocystum. Int. J. Appl. Innov. Eng. Manag. 3: 39–45.
Sydney, E. B., A. C. Novak, J. C. de Carvalho and C. R. Soccol. 2014. Respirometric balance and carbon fixation of
industrially important algae. In Biofuels from algae. Elsevier. 67–84.
Tchounwou, P. B., C. G. Yedjou, A. K. Patlolla and D. J. Sutton. 2012. Heavy metal toxicity and the
environment. Molecular, Clinical and Environmental Toxicology. 133–164.
222 Bioremediation for Sustainable Environmental Cleanup
Toranzo, R., G. Ferraro, M. V. Beligni, G. L. Perez, D. Castiglioni, D. Pasquevich and C. Bagnato. 2020. Natural and
acquired mechanisms of tolerance to Cr in a S. dimorphus strain. Algal Res. 52: 102100.
Ubando, A. T., A. D. M. Africa, M. C. Maniquiz-Redillas, A. B. Culaba, W. H. Chen and J. S. Chang. 2021. Microalgal
biosorption of heavy metals: a comprehensive bibliometric review. J. Hazard. Mater. 402: 123431.
Uzunoğlu, D., N. Gürel, N. Özkaya and A. Özer. 2014. The single batch biosorption of copper (II) ions on
S. acinarum. Desalin. Water Treat. 52(7-9): 1514–1523.
Vijayaraghavan, K. and Y. S. Yun. 2008. Bacterial biosorbents and biosorption. Biotechnol. Adv. 26(3): 266–291.
Viswanath, B. and F. Bux. 2012. Biodiesel production potential of wastewater microalgae Chlorella sp. Under
photoautotrophic and Heterotrophic growth conditions. Br. J. Eng. Technol. 1: 251–264.
Volesky, B. and Z. R. Holan. 1995. Biosorption of heavy metals. Biotechnol. Progress. 11(3): 235–250.
Volesky, B. 2003a. Biosorption process stimulation tools. Hydrometall.71(1-2): 179–190.
Volesky, B.2003b. Sorption and biosorption. St. Lambert, Quebee: BV Soebex.
Waldichuk, M. 1974. Some biological concerns in heavy metals pollution. Pollution and physiology of marine
Organisms. 1: 1–59.
Yeong, T. K., K. Jiao, X. Zeng, L. Lin, S. Pan and M. K. Danquah. 2018. Microalgae for biobutanol production–
Techno. Evaluation and value proposition. Algal Res. 31: 367–376.
Zeitoun, M. Moustafa and E. E. Mehana. 2014. Impact of water pollution with heavy metals on fish health: overview
and updates. Global Vet. 12: 219–231.
Zeraatkar, A. K., H. Ahmadzadeh, A. F. Talebi, N. R. Moheimani and M. P. McHenry. 2016. Potential use of algae for
heavy metal bioremed., a critical review. J. Environ. Manag. 181: 817–831.
Zhang, L. and M. H. Wong. 2007. Environmental Hg contamination in China: sources and impacts. Environ. Int.
33(1): 108–121.
Zhou, Q., N. Yang, Y. Li, B. Ren, X. Ding, H. Bian and X. Yao. 2020. Total concentrations and sources of heavy
metal pollution in global river and lake water bodies from 1972 to 2017. Global Ecology and Conservation.
22: e00925.
Chapter 13
Bioremediation of Heavy
Metals from Aquatic
Environments
Fariha Latif,1,* Shahena Perveen,1 Sana Aziz,2
Rehana Iqbal,1 and Muhammad Mudassar Shahzad 3
13.1 Introduction
The severe contamination of freshwater resources, particularly in developing countries, is
causing serious health hazards and environmental issues. Though aquatic pollution is not a novel
phenomenon, but the pace of the industrial revolution and urbanization have aggravated its negative
impacts on the aquatic environment. Predominant anthropogenic sources of aquatic contamination
include mining operations, untreated industrial effluents, domestic sewage, waste dump leachates
and combustion emissions. These sources are adding a number of pollutants including metallic ions,
polycyclic aromatic hydrocarbons, dioxins, polychlorinated biphenyls and other xenobiotics in the
natural aquatic ecosystems. Metallic ions toxicity affects the physiology and ecology of aquatic
organisms due to their specific properties of long persistence, bioaccumulation and biomagnification
in the food chain. Currently, all areas of the environment are being polluted by several chemical and
biological pollutants and the aquatic ecosystems are serving as a major repository for them. Among
these pollutants, metallic ions pollution is widespread and has become an issue of great concern
as metallic ions are indiscriminately discharged into freshwaters, deteriorating their quality and
ultimately affecting the inhabitant fish fauna.
1
Institute of Zoology, Bahauddin Zakariya University, Multan, Pakistan.
2
Department of Zoology, Wildlife and Fisheries, University of Agriculture, Faisalabad, Pakistan.
3
Department of Zoology, Division of Science & Technology, University of Education, Lahore, Pakistan.
* Corresponding author: [email protected]
224 Bioremediation for Sustainable Environmental Cleanup
Water and soil are now becoming polluted due to the release of garbage. Untreated wastewater
is added to water, and it contains various metals that create pollution (Rudakiya and Pawar 2013).
Heavy metals have a negative effect on the food web and food chain. They can also impact numerous
activities of cells (Vardhan et al. 2019).
13.2 Bioremediation
The method in which plants, microorganisms, bacteria and fungi are used to eliminate pollutants
from the environment is called bioremediation. The purpose of “bioremediation” methods is to
decrease the level of the contaminant to a minimum and within safe limits viz. within the range
set by different environmental protection agencies (Pointing 2001). Bioremediation can also be
increased by adding different materials such as nitrogen and carbon sources (Rudakiya et al. 2019a).
The fungi and bacteria can augment the bioremediation process (Zaki et al. 2014).
The bioremediation process can reduce negative impacts on the natural environment. It is a less
costly process. The total remediation of contaminants is attainable on-site without the requirement
of excavation (Vidali 2001). It needs a small amount of energy and protects the soil. It can be made
more reliable to the community (Zhang and Qiao 2002). Aerobic soil is essential for the fungal
system to work. There are many drawbacks to the bioremediation process. This method can be used
for those compounds that are biodegradable. The by-products of the bioremediation process are
less hazardous than the original compounds. These processes are highly specified and complex. It
requires a longer duration for their work (Vidali 2001).
The bioremediation of heavy metals is done by using the following processes (Figure 13.2):
13.2.1 Biosorption
13.2.2 Bioaccumulation
13.2.3 Bioleaching
13.2.4 Biotransformation
13.2.5 Biomineralization
Biotransformation
Bioaccumulation Biomineralization
Bioremediation
Biosorption Bioleaching
13.2.1 Biosorption
The capacity of biological materials to concentrate heavy metals from wastewater via metabolically
mediated or physico-chemical pathways is called biosorption (Fourest and Roux 1992). The
biological material can bind heavy metals from an aqueous solution. It has the characteristics of
some substances to concentrate and bind some ions from metal-containing water (Volesky 2007).
The best biosorbents that adsorb metals are bacteria, fungi, yeasts and algae. Biosorption is an
alternative rapid method for the removal of pollutants by using non-active and nonliving microbes.
The advantage of this process are as follows:
No additional nutrient requirement
biosorbent can be regenerated
metal recovery
low cost
high efficiency
decreasing amount of biological sludge
The process of biosorption is affected by various physico-chemical parameters such as
temperature, pH, metal ion’s initial concentration, biosorbent concentration and speed of mixing of
biosorbent and solution. The entire biomass of biosorbent can be treated physically and chemically
before use and can be made economical by recycling and reusing the biosorbent material after
removing the metals. Various bioreactors can be used in biosorption for the removal of metal ions
from large volumes of water or effluents (Kanamarlapudi et al. 2018). When the physico-chemical
interaction of metal ions occurs with cellular compounds heavy metals can be adsorbed.
13.2.2 Bioaccumulation
This is an active metabolic process that relies on the import-storage system. Using transporter
proteins, the system moves heavy metal ions through the lipid bilayer and into the cytoplasm or
intracellular regions. The heavy metals that these entities sequester might exist in particulate forms,
insoluble forms and their by-products. The metal ions are sequestered by metal-binding entities such
as proteins and peptide ligands. The process in which ions of metals accumulate in the living cell
is called bioaccumulation. It is a complex and irreversible process that depends on cell metabolism
and is mainly done by microorganism biomass cultivation. The metabolic process starts in the
organism that activates the intracellular transport system (Srichandan et al. 2014, Gu et al. 2018).
A few examples include; the yeast Pichia stipites can bio-accumulate Cr3+ and Cu2+ with the highest
absorbance ability of 9.10 mg g–1 and 15.85 mg g–1, respectively (Ojha et al. 2015). Aspergillus
niger can eliminate Pb and Cu with the greatest uptake of 34.4 mg g–1 and 15.6 mg g–1, respectively
(Srichandan et al. 2015).
13.2.3 Bioleaching
The bioleaching process is also called bio-hydrometallurgy or microbial leaching. The process in
which solubilization of metals occurs from some ores that are not soluble is called bioleaching.
Microorganisms are also used in this process as reducing agents (Mishra et al. 2005). These methods
include membrane separation, solvent extraction, adsorption, ion exchange, selective precipitation
and electrowinning.
Microbial bioleaching is used for the solubilization of various metals viz. cobalt, nickel copper,
zinc and uranium (Rohwerder et al. 2003). For example, in the bioleaching process of arsenopyrite,
the sulfides of arsenic and iron are dissolved. The endured gold is recuperated by the procedure
known as cyanidation.
Bioremediation of Heavy Metals from Aquatic Environments 227
The natural process of bioleaching has its limitations therefore, researchers have broadened
this process to treat solid wastes artificially to remove or solubilize metals. These solid wastes are
released from different industries and mining processes causing various health hazards to animal
diversity as well as human health. Different microbes are used in the bioleaching process. Examples
of some microbes are archaea, fungi and acidophilic bacteria (Brandl et al. 2001, Natarajan and
Ting 2014). All three types of acidophilic bacteria such as mesophiles, moderate thermophiles and
thermophiles are used in this process. Fungi can also be used in this process. Some of the fungi that
are used in the bioleaching process include Penicillium chrysogenum, Penicillium simplicissimum,
Aspergillus niger and Aspergillus flavus. The fungal bioleaching process requires a pH from 3.0 to
7.0 and a temperature ranging from 25 to 35°C.
13.2.4 Biotransformation
This is the method by which a chemical compound’s structure is changed, resulting in the production
of a molecule with comparatively higher polarity. In other words, through the process of metal-
microbe interaction, metal and organic molecules are changed from a harmful state to a form that
is substantially less toxic. This technique basically enables microbes to acclimatize to changing
environments.
Microorganisms control trace element transformation (microbial or biotransformation) through
a number of mechanisms such as oxidation, reduction, methylation, demethylation, complex
formation and biosorption. Microbial transformation is important in the behavior and fate of toxic
elements in soils and sediments, particularly Arsenic (As), Chromium (Cr), mercury (Hg), and
Selenium (Se). Biotransformation processes can change the speciation and redox state of these
elements, controlling their solubility and mobility (Kunhikrishnan et al. 2017). These processes are
critical for trace element bioavailability, mobility, ecotoxicity and environmental health. Microbial
cells have a high surface-to-volume ratio, a rapid rate of growth, as well as metabolic activity, and
is simple to maintain sterile conditions for microbes. They are thus ideal organisms for the process
of biotransformation. Condensation, hydrolysis, the creation of new carbon bonds, isomerization,
the addition of functional groups, oxidation, reduction and methylation are all methods that can be
used to carry out this process. These processes might cause metals to volatilize, thus decreasing their
ability (Tayang and Songachan 2021).
13.2.5 Biomineralization
A natural process of mineral production is the biomineralization of heavy metals. Minerals like
phosphates, oxides, sulfates, silicates and carbonates are naturally synthesized in this process, which
involves a variety of mechanisms in living things. Mineral production depends on the presence
of highly variable and reactive surfaces, such as cell walls and extra organic layers with varying
levels of hydration, content and structure. Additionally, there are organic ligands that deprotonate
and impart a net negative charge on the microbial surface as pH rises, including amine, carboxyl,
hydroxyl, phosphoryl and sulfur.
Positively charged potential hazardous metals precipitate unevenly into more stable and
compact mineral products. Phosphate precipitation, carbonate precipitation, oxalate precipitation
and complexation can all result in the immobilization or complexation of metals (Tayang and
Songachan 2021).
the unsolvable form of metal (Rudakiya et al. 2019). The following kinds of bioremediations are
used to eliminate toxic heavy metals, metalloids and non-metals from different areas:
13.3.1 Bacterial remediation
13.3.2 Fungal remediation
13.3.3 Algal remediation
13.3.4 Phyto remediation
and the second method involves the absorption of metals by the metabolism called bioaccumulation
(Kadukova and Vircikova 2005). Some of the studies that are based on the biosorption process are
detailed in Table 13.1.
These methods are surface binding, the complexion by various functional groups including
hydroxyl, phosphate, carboxyl, and amine groups and ion exchange reaction (Wang and Chen
2009). The fungus cell wall makes a connection with ions of metals (Figure 13.3). The extracellular
methods use cell wall binding, precipitation and chelation (Bellion et al. 2006). The recycling of
fungi in bioremediation is an expensive process (Tsezos 1984). Living cells of fungi are killed by
various methods like chemical drying, mechanical drying, autoclaving and vacuum drying.
Factors Affecting the Fungal Biosorption: The removal of ‘heavy metals’ by fungi depends on
various environmental factors and types of metals (Hassen et al. 1998). Different parameters such
as oxygen level and glucose affect the absorption of metals (Javaid et al. 2011). The parameters like
concentration, contact time, initial metal ion concentration, temperature, pH and biomass have been
noted to affect the fungal biosorption process (Kapoor and Viraraghavan 1995).
Fungal Bioleaching: Fungal bioleaching has been used in mining areas and for the treatment of
waste material containing metals (Bosecker 2001). Fungi can also be used for sorting of heavy
metals from low-grade ores (Chaudhary et al. 2014). Penicillium verruculosum and Aspergillus
were used to solubilize iron Penicillium and Aspergillus leached Ni, Fe and Co (Valix et al. 2001).
Fungal Bio-immobilization: The interaction between metals and fungus causes a notable decrease
in the mobility of heavy metals. This immobility has been achieved by different methods such as
bio-precipitation, bio-reduction/bio-oxidation, biosorption and bioaccumulation. The fungus,
Rhizopus arrhizus can be used for the elimination of Cu2+, Fe3+ and Cd2+ by the fungal
bio-immobilization process (Lewis and Kiff 1998).
Fungal Biomineralization: Biologically Controlled Mineralization (BCM) and Biologically
Induced Mineralization (BIM) are types of biomineralization. In BCM, the microorganisms are
used to control the growth and nucleation of the biominerals, however in BIM organisms modify
the native environment to produce optimum conditions for the mineral’s precipitation (Gadd 2010).
230 Bioremediation for Sustainable Environmental Cleanup
Biomineralization,
sulfite,
are alginate and cellulose. Its cell wall also provides various functional groups including imidazole,
phosphate, sulfonate, thiol, amino, carboxyl and hydroxyl. These functional groups can bind with
heavy metals (Priatni et al. 2017). They also contain a large amount of alcohol and carboxyl groups
that attract the cations and anions of heavy metals (Pradhan et al. 2019).
The spectroscopy including NMR (Nuclear Magnetic Resonance) and FTIR (Fourier transform
infrared) can be used to find out different functional groups of microalgae. Various methods can be
used for the adsorption of heavy metals on the outer surface of microalgae. These methods include
the ionic exchange of heavy metal ions with cell wall cation. The heavy metals are moved into the
cytoplasm by the process of diffusion (Ibuot et al. 2017, Pradhan et al. 2019).
Algal Bioremediation of a few metals
Arsenic (As) is a toxic heavy metal. Its pollution is caused by human activities including medical
use, smelting and mining (Arora et al. 2017, Singh et al. 2016). Microalgae are used to reduce the
pollution of arsenic by the complex formation with glutathione (Papry et al. 2019).
Cadmium (Cd) is poisonous heavy metal. It is released into the surroundings by human activities
including color pigments, tanning industries and pesticide use (Abinandan et al. 2019). The dead
microalgae can be used for bioremediation of Cd by surface binding mechanism.
Chromium (Cr) is used in ink, paint and dyes (Gupta and Rastogi 2008). Microalgae can be
used for the removal of chromium by using the mechanism of biosorption. Navicula pelliculosa and
Phaeodactylum tricornutum have been used for this purpose (Hedayatkhah et al. 2018).
13.3.4 Phytoremediation
The use of living plants to remove heavy metals and metalloids from water, soil and the air is called
phytoremediation. The types of phytoremediation are:
• Phytoextraction
• Phytostabilization
232 Bioremediation for Sustainable Environmental Cleanup
• Phytofiltration
• Phytovolatilization
• Phytodegradation
Phytoextraction
It is also called phytoabsorption or phtyoaccumulation and phytosequestration. The process which
involves the elimination of toxic contaminants such as heavy metals and metalloids from water
and soil by using the roots of plants is known as phytoextraction (Ghosh and Singh 2005). These
pollutants can be moved and collected in shoots (Muthusaravanan et al. 2018). The process in which
contaminants can be gathered in shoots is a very important procedure as if the material is gathered
in roots, then harvesting of biomass of roots is not possible (Halim et al. 2003, McIntyre 2003).
If phytoextraction is done continuously then a very large number of pollutants can be collected in
plants (Sarwar et al. 2017).
The following steps are involved in the phytoextraction process:
• The first step involves the mobilization of contaminants in the rhizosphere.
• The roots can absorb toxic contaminants like heavy metals and metalloids (Memon and Schröder
2009).
• The pollutants are translocated toward the airy parts.
• The contaminants are separated from the tissues of plants (Ali et al. 2013).
For the phytoextractions, the hyperaccumulator plants are used that must have the following
properties: the growth rate of plants should be a high, large biomass (Khalid et al. 2017), tolerance
to a wide variety of contaminants, the branched root system, adjustment to native environmental
situation, immunity against pest and pathogen, high crop production and a higher ability to
accumulate pollutants (Mahar et al. 2016, Sarwar et al. 2017).
The efficacy of the phytoextraction process depends upon the following three facts viz. the
climatic condition (Bhargava et al. 2012), weather and the growth rate of plant roots.
This process can be improved by using various agents such as nitrilotriacetic acid, amino
polycarboxylic acid, ethylenediaminetetraacetic acid, nitrilotriacetic acid and citric acid. This
process is used for the elimination of contaminants from the contaminated areas.
There are also many disadvantages to this process. As it requires specific conditions that are
necessary for the plant’s tolerance to contaminants and their growth. It also requires a longer time
to completely remediate the areas.
Phytostabilization
It is also called phytoimmobilization. The process in which plants can be used to reduce the
bioavailability or mobility of contaminants either by ceasing the leaching to the underground water
table (Sarwar et al. 2017) or its entrance into the food chain through a process such as the production
of compounds that are not soluble in the roots zone or adsorption by the roots of plants (Khan et al.
2019). The properties of this process are:
• The reduction of contaminants in the polluted region by assimilation and adsorption on the
roots and precipitation by roots.
• The deployment of the plant’s roots to avoid pollutants movements by water, wind, draining
and soil dispersal (USEPA 1999).
The goal of this process is the stabilization of contaminants rather than their elimination. It
decreases their risk to humans and nature (Bolan et al. 2011). Phytostabilization not only reduces the
concentration of contaminants, but also lessens the contaminant of the surrounding region (Khalid
et al. 2017).
Bioremediation of Heavy Metals from Aquatic Environments 233
13.4 Conclusion
Bioremediation is an effective and environment-friendly technique to biologically degrade various
pollutants. In the bioremediation process, many microorganisms play a major role such as bacteria,
fungi, algae and other plants. Furthermore, fungus microorganisms can effectively degrade many
toxic environmental pollutants. Phytoremediation represents an emerging technology through which
plants can be used to remove pollution from the soil, water and other environments. Bioremediation
requires less effort, is less labor intensive, cheap, eco-friendly, sustainable and relatively easy to
implement. Various bioremediation methods are used to convert toxic heavy metals into non-toxic
or environmentally friendly products. Biosorption, bioaccumulation, bioleaching, biotransformation
234 Bioremediation for Sustainable Environmental Cleanup
and biomineralization are examples of these processes. Furthermore, because there is no acceptable
endpoint, evaluating the performance of bioremediation may be difficult. More research is needed
to develop bioremediation technologies and find more biological solutions for the bioremediation of
heavy metal contamination from various environmental systems.
References
Abinandan, S., S. R. Subashchandrabose, K. Venkateswarlu, I. A. Perera and M. Megharaj. 2019. Acid-tolerant
microalgae can withstand higher concentrations of invasive cadmium and produce sustainable biomass and
biodiesel at pH 3.5. Bioresour. Technol. 281: 469–473.
Acosta-Rodríguez, I., J. F. Cárdenas-González, A. S. Rodríguez Pérez, J. T. Oviedo and V. M. Martínez-Juárez. 2018.
Bioremoval of different heavy metals by the resistant fungal strain Aspergillus niger. Bioinorg. Chem. 2018.
1–7.
Ali, H., E. Khan and M. A. Sajad. 2013. Phytoremediation of heavy metals-concepts and applications. Chemosphere.
91: 869–881.
Arora, N., K. Gulati, A. Patel, P. A. Pruthi, K. M. Poluri and V. Pruthi. 2017. A hybrid approach integrating arsenic
detoxification with biodiesel production using oleaginous microalgae. Algal Res. 24: 29–39.
Bai, R. S. and T. E. Abraham. 2001. Biosorption of Cr (VI) from aqueous solution by Rhizopus nigricans. Bioresour.
Technol. 79: 73–81.
Balaji, S., T. Kalaivani, M. Shalini, M. Gopalakrishnan, M. A. Rashith and C. Rajasekaran. 2016. Sorption sites of
microalgae possess metal binding ability towards Cr (VI) from tannery effluents-a kinetic and characterization
study. Desalin. Water Treat. 31: 14518–14529.
Baldrian, P. 2003. Interactions of heavy metals with white-rot fungi. Enzym. Microb. Technol. 32: 78–91.
Bellion, M., M. Courbot, C. Jacob, D. Blaudez and M. Chalot. 2006. Extracellular and cellular mechanisms sustaining
metal tolerance in ectomycorrhizal fungi. FEMS Microbiol. Letters. 254(2): 173–181.
Bhargava, A., F. F. Carmona, M. Bhargava and S. Srivastava. 2012. Approaches for enhanced phytoextraction of
heavy metals. J. Environ. Manag. 105: 103–120.
Birungi, Z. S. and E. M. N. Chirwa. 2015. The adsorption potential and recovery of thallium using green micro-algae
from eutrophic water sources. J. Hazard. Mater. 299: 67–77.
Bolan, N. S., J. H. Park, B. Robinson, R. Naidu and K. Y. Huh. 2011. Phytostabilization. A green approach to
contaminant containment. Adv. Agron. 112: 145–204.
Bosecker, K. 2001. Microbial leaching in environmental clean-up program. Hydrometallurgy. 59: 245–248.
Brandl, H., R. Bosshard and M. Wegmann. 2001. Computer-munching microbes: metal leaching from electronic
scrap by bacteria and fungi. Hydrometallurgy. 59: 319.
Bulak, P., A. Walkiewicz and M. Brzeziska. 2014. Plant growth regulators-assisted phytoextraction. Biol. Plant.
58: 1–8.
Cameron, H., M. T. Mata and C. Riquelme. 2018. The effect of heavy metals on the viability of Tetraselmis marina
AC16-MESO and an evaluation of the potential use of this microalga in bioremediation. Peer J. 6: 5295.
Chaudhary, N., T. Banerjee and N. Ibrahim. 2014. Fungal leaching of iron ore: isolation, characterization and
bioleaching studies of Penicillium verruculosum. Am. J. Microbiol. 50: 27–32.
Couto, S. R., M. A. Sanroman, D. Hoefer and G. M. Gubitz. 2004. Stainless steel: a novel career for the immobilization
of the white rot fungus Trametes hirsute for decolorization of textile dyes. Bioresour. Technol. 95: 67–72.
Das, D., A. Chakraborty, S. Bhar, M. Sudarshan and S. C. Santra. 2013. Gamma irradiation in modulating cadmium
bioremediation potential of Aspergillus sp. IOSR J. Environ. Sci. Toxicol. Food Technol. 3: 51–55.
Dhanam, S. 2017. Strategies of bioremediation of heavy metal pollutants toward sustainable agriculture.
pp. 349–358. In: A. Dhanarajan [Ed.]. Sustainable Agriculture Towards Food Security. Springer Nature,
Singapore.
Dhankhar, R. and A. Hooda. 2011. Fungal biosorption—an alternative to meet the challenges of heavy metal pollution
in aqueous solutions. Environ. Technol. 32: 467–491.
Fourest, E. and J. Roux. 1992. Heavy metal biosorption by fungal mycelial by-products: mechanisms and influence
of pH. Appl. Microbiol. Biotechnol. 37: 399–403.
Gadd, G. M. 2010. Metals, minerals and microbes: geomicrobiology and bioremediation. Microbiology. 156: 609.
Ghosh, M. and S. P. Singh. 2005. A review on phytoremediation of heavy metals and utilization of it’s by products.
Asian J. Energy Environ. 6(4): 18.
Gomes, M. A. C., R. A. Hauser-Davis, A. N. De Souza and A. P. Vitória. 2016. Metal phytoremediation: general
strategies, genetically modified plants and applications in metal nanoparticle contamination. Ecotoxicol.
Environ. Saf. 134: 133–147.
Bioremediation of Heavy Metals from Aquatic Environments 235
Maheshwari, S. and A. G. Murugesan. 2009. Remediation of arsenic in soil by Aspergillus nidulans isolated from an
arsenic-contaminated site. Environ. Technol. 30(9): 921–926.
McIntyre, T. 2003. Phytoremediation of heavy metals from soils. Adv. Biochem. Eng. Biotechnol. 78: 97–123.
Memon, A. R. and P. Schröder. 2009. Implications of metal accumulation mechanisms to phytoremediation. Environ.
Sci. Pollut. Res. 16: 162–175.
Mishra, D., D. J. Kim, J. G. Ahn and Y. H. Rhee. 2005. Bioleaching: a microbial process of metal recovery; a review.
Met. Mater. Int. 11: 249.
Muthusaravanan, S., N. Sivarajasekar, J. S. Vivek, T. Paramasivan, M. U. Naushad and J. Prakashmaran. 2018.
Phytoremediation of heavy metals: mechanisms, methods and enhancements. Environ. Chem. Lett. 16: 1339–
1359.
Natarajan, G. and Y. P. Ting. 2014. Pretreatment of e-waste and mutation of alkali-tolerant cyanogenic bacteria
promote gold biorecovery. Bioresour. Technol. 152: 80.
Ojha, S. K., S. Mishra, S., Kumar, S. S. Mohanty, B. Sarkar, M. Singh and G. R. Chaudhury. 2015. Performance
evaluation of vinasse treatment plant integrated with physico-chemical methods. J. Environ. Biol. 36: 1269.
Papry, R. I., K. Ishii, M. A. A. Mamun, S. Miah, K. Naito, A. S. Mashio, T. Maki and H. Hasega. 2019. Arsenic
biotransformation potential of six marine diatom species: effect of temperature and salinity. Sci. Rep. 9(1):
10226.
Pointing, S. 2001. Feasibility of bioremediation by white-rot fungi. Appl. Microbiol. Biotechnol. 57: 20–33.
Pradhan, D., L. B. Sukla, B. B. Mishra and N. Devi. 2019. Biosorption for removal of hexavalent chromium using
microalgae Scenedesmus sp. J. Cleaner Prod. 209: 617–629.
Priatni, S., D. Ratnaningrum, S. Warya and E. Audina. 2017. Phycobiliproteins production and heavy metals reduction
ability of Porphyridium sp. IOP Conf. Series: Earth Environ. Sci. 60.
Roane, T. M., I. L. Pepper and T. J. Gentry. 2015. Microorganisms and metal pollutants. pp. 415–439. In: I. L. Pepper,
C. P. Gerba and T. J. Gentry [eds.]. Environmental Microbiology. Academic Press, San Diego, CA.
Rohwerder, T., T. Gehrke, K. Kinzler and W. Sand. 2003. Bioleaching review part A. Appl. Microbiol. Biotechnol.
63: 239.
Rudakiya, D., Y. Patel, U. Chhaya and A. Gupte. 2019. Carbon nanotubes in agriculture: production, potential, and
prospects. pp. 121–130. In: D. G. Panpette and Y. K. Jhala [eds.]. Nanotechnology for Agriculture. Springer,
Singapore.
Rudakiya, D. M. and K. S. Pawar. 2013. Evaluation of remediation in heavy metal tolerance and removal by
Comamonas acidovorans MTCC 3364. IOSR J. Environ. Sci. Toxicol. Food Technol. 5: 26–32.
Rudakiya, D. M. and A. Gupte. 2017. Degradation of hardwoods by treatment of white rot fungi and its pyrolysis
kinetics studies. Int Biodeterior Biodegradation. 120: 21–35.
Rudakiya, D. M. and K. Pawar. 2017. Bactericidal potential of silver nanoparticles synthesized using cellfree extract
of Comamonas acidovorans: in vitro and in silico approaches. 3 Biotech. 7(2): 9.
Rudakiya, D. M. 2018. Metal tolerance assisted antibiotic susceptibility profiling in Comamonas acidovorans.
Biometals. 31(1): 1–5.
Rudakiya, D. M., A. Tripathi, S. Gupte and A. Gupte. 2019. Fungal bioremediation: a step towards cleaner
environment. pp. 229–249.. In: T. Satyanarayana, S. K. Deshmukh and M. V. Deshpande [eds.]. Advancing
Frontiers in Mycology & Mycotechnology. Springer, Singapore.
Rudakiya, D. M., D. H. Patel and A. Gupte. 2020. Exploiting the potential of metal and solvent tolerant laccase from
Tricholoma giganteum AGDR1 for the removal of pesticides. Int. J. Biol. Macromol. 144: 586–595.
Sarwar, N., M. Imran, M. R. Shaheen, W. Ishaque, M. A. Kamran, A. Matloob, A. Rehim and S. Hussain. 2017.
Phytoremediation strategies for soils contaminated with heavy metals: modifications and future perspectives.
Chemosphere. 171: 710–721.
Satyanarayana, T., S. K. Deshmukh and M. V. Deshpande. 2019. Advancing frontiers in mycology and mycotechnology:
basic and applied aspects of fungi. Springer Nature, Singapore.
Say, R., N. Yilmaz and A. Denizli. 2004. Removal of chromium (VI) ions from synthetic solutions by the fungus
Penicillium purpurogenum. Eng. Life Sci. 4: 276–280.
Shah, D., D. M. Rudakiya, V. Iyer and A. Gupte. 2018. Simultaneous removal of hazardous contaminants using
polyvinyl alcohol coated Phanerochaete chrysosporium. Int. J. Agri. Environ. Biotechnol. 11(2): 235–241.
Siegel, S. M., B. Z. Siegel and K. E. Clark. 1983. Biocorrosion: solubilization and accumulation of metals by fungi.
Water Air Soil Pollut. 19: 229–236.
Singh, N. K., A. S. Raghubanshi, A. K. Upadhyay and U. N. Rai. 2016. Arsenic and other heavy metal accumulation
in plants and algae growing naturally in contaminated area of West Bengal, India. Ecotoxicol. Environ. Saf.
130: 224–233.
Bioremediation of Heavy Metals from Aquatic Environments 237
Srichandan, H., A. Pathak, S. Singh, K. Blight, D. J. Kim and S. W. Lee. 2014. Sequential leaching of metals from
spent refinery catalyst in bioleaching-bioleaching and bioleaching-chemical leaching reactor: comparative
study. Hydrometallurgy. 150: 130.
Srichandan, H., S. Singh, K. Blight, A. Pathak, D. J. Kim, S. Lee and S. W. Lee. 2015. An integrated sequential
biological leaching process for enhanced recovery of metals from decoked spent petroleum refinery catalyst:
a comparative study. International Journal of Mineral Processing 134: 66–73.
Tayang, A. and L. S. Songachan. 2021. Microbial bioremediation of heavy metals. Current Science. 120(6): 00113891.
Tobin, J. M., D. G. Cooper and R. J. Neufeld. 1984. Uptake of metal ions by Rhizopus arrhizus biomass. Appl.
Environ. Microbial. 47: 821–824.
Townslcy, C. C., I. S. Ross and A. S. Atkins. 1986. Biorecovery of metallic residues from various industrial effluents
using filamentous fungi. pp. 279–289. In: R. W. Lawrence, B. RMR and H. G. Enner [Eds.]. Fundamental and
Applied Biohydrometallurgy. Elsevier, Amsterdam.
Tsezos, M. 1984. Recovery of uranium from biological adsorbents-desorption equilibrium. Biotechnol. Bioeng.
26: 973–981.
USEPA. 1999. Report on Bioavailability of Chemical Wastes With Respect to the Potential for Soil Remediation.
T28006: QT-DC-99-003260.
Valix, M., F. Usai and R. Malik. 2001. Fungal bio-leaching of low-grade laterite ores. Miner Eng. 14: 197–203.
Vardhan, K. H., P. S. Kumar and R. C. Panda. 2019. A review on heavy metal pollution, toxicity and remedial
measures: current trends and future perspectives. J. Mol. Liq. 111: 197.
Veglio, F. and F. Beolchini. 1997. Removal of metals by biosorption: a review. Hydrometallurgy. 44: 301–316.
Vidali, M. 2001. Bioremediation. An overview. Pure Appl Chem. 73(7):1163–1172.
Volesky, B. 2007. Biosorption and me. Water Res. 41(18): 4017–4029.
Wang, J. and C. Chen. 2009. Biosorbents for heavy metals removal and their future. Biotechnol. Adv. 27: 195–2.
Zaki, M. S., N. El-Battrawy, A. M. Hammam and S. I. Shalaby. 2014. Aquatic bioremediation of metals. Life Sci. J.
11(4): 1394.
Zhang, J. L. and C. L. Qiao. 2002. Novel approaches for remediation of pesticide pollutants. Int. J. Environ. Pollut.
18(5): 423–433.
Chapter 14
Phytoremediation of
Wastewater Discharged
from Paper and Pulp, Textile
and Dairy Industries using
Water Hyacinth
(Eichhornia crassipes)
A Sustainable Approach
Apurba Koley,1 Anudeb Ghosh,1 Sandipan Banerjee,2
Nitu Gupta,3 Richik Ghosh Thakur,1
Binoy Kumar Show,1 Shibani Chaudhury,1
Amit Kumar Hazra,4 Andrew B. Ross,5
Gaurav Nahar 6 and Srinivasan Balachandran1,*
14.1 Introduction
The presence of toxic heavy metals in water bodies like rivers, lakes and reservoirs affect local
inhabitants, animals and other living organisms depending on the water or source (Rai et al. 2002).
The concentration of these metals above the threshold limit or optimal concentration can cause
negative impacts on plant systems due to oxidative stress and inhibited cytoplasmic enzymes
(Assche and Clijsters 1990). Intake of toxic metals causes serious health and growth-related issues
1
Bioenergy Laboratory, Department of Environmental Studies, Siksha-Bhavana, Visva-Bharati.
2
Mycology and Plant Pathology Laboratory, Department of Botany, Visva-Bharati University, Santiniketan-731235, India.
3
Department of Environmental Science, Tezpur University, Napaam, Tezpur, Assam - 784208, India.
4
Socio-Energy Lab, Department of Lifelong Learning and Extension, Visva-Bharati, Santiniketan.
5
School of Chemical and Process Engineering, University of Leeds, Leeds, LS2 9JT, United Kingdom.
6
Defiant Renewables Pvt Ltd, 1st Floor Kant Helix, Bhoir Colony, Chinchwad, Pune-411033, India.
* Corresponding author: [email protected]
ORCID ID: 0000-0003-4247-408X
Phytoremediation of Industrial Wastewater using Water Hyacinth 239
through food-chain magnification (Rai and Tripathi 2008, Pramanik et al. 2021). Different removal
methods like electrolysis, ion exchange, precipitation, absorption and reverse osmosis are used
to remove heavy metals. These methods are quite expensive and are comparatively ineffective,
producing a large amount of waste that is difficult to dispose of. The development of cost-effective
and alternative methods for the removal of toxic substances from wastewater is a necessity (Rai and
Tripathi 2007). The use of living plants to remove toxic substances from the water and soils could
be one of the alternatives and viable processes (Rai 2009). Plants have the capability to accumulate
metals (Co, Ca, Fe, Cu, Ni, Mo, Mg, Mn, Na, Ni, Se, V and Zn) and non-essential metals (Al, As, Au,
Cd, Cr, Hg, Pb, Pd, Pt, Sb, Te, Tl and U) under different concentration (Djingova and Kuleff 2000).
Three different patterns have been observed during the uptake of heavy metals by plants (i). True
exclusion (metals are limited during entering), (ii). Shoot exclusion (root accumulates the metals
but restricts translocation to the shoot system), and, (iii). Accumulation (plant parts concentrate the
metals) (Zavoda et al. 2001, Kamal et al. 2004).
Textile wastewaters are a serious polluting threat as they contaminate the environment; without
proper treatment, they lead to irreversible persistent changes in the environment. Textile wastes
include dyes and have high Biological Oxygen Demand (BOD), Chemical Oxygen Demand (COD),
Total Suspended Solids (TSS), Total Nitrogen (TN), heavy metals, phosphates and greases, etc.
Dyes are well-known toxins which can directly affect the aquatic ecosystem (Ekanayake et al. 2021).
The pulp and paper industries are one of the major causes of aquatic pollution in India (Saadia
and Ashfaq 2010, Singh et al. 2022). The major pollutants in paper and pulp industries are high in
TSS, Total Dissolved Solids (TDS), Total Nitrogen (TN), adsorbable organic halides (AOX), heavy
metals, as well as have high COD. A large amount of this contaminated wastewater is generated and
discharged during the paper-making process, which negatively impacts wild and human life. The
impurities should be removed or minimized before being discharged into the environment (Singh
et al. 2022, Ashrafi et al. 2015).
Similar to the textile and paper industries, the Indian dairy industry generates wastewater which
is a major cause of concern. Six to seven liters of effluent is generated while processing one liter
of milk (Porwal 2015). High levels of BOD, COD, hardness, along with ions like K+, Na+ and Cu2+
ions are detected in water close to the industry due to inappropriate disposal of effluents in rivers
(Kaur et al. 2018). To reduce the release of these environmental pollutants from such industries,
many researchers are trying to develop sustainable, cost-effective treatment practises for cleaning
the industrial effluents.
Phytoremediation is one such technique being researched to clean up wastewater generated from
industrial and domestic discharge for consumption of fresh water. Eichhornia crassipes, i.e., Water
Hyacinth (WH) is one of the promising plants popularly tested for phytoremediation of various
wastewater streams (Koley et al. 2022). This aquatic macrophyte is a large-leaved, noxious, free-
floating aquatic weed that belongs to the family Pontederiaceae. WH is a free-floating angiosperm
having roots, short to long stems, broad, glossy leaves and lavender flowers, usually found in
shallow water. Due to its rapid reproduction rate, it covers the water surface by forming a thick mat.
Owing to its fast-growing and high nutrient uptake ability, WH is used in constructed wetlands to
uptake pollutants from different wastewater sources. Even though there are many sustainable uses
of WH, most research has been done on how to use it to clean up nutrients and contaminants, like
heavy metals, from industrial and urban wastewater (Koley et al. 2022, Sinha et al. 2021, Qin et al.
2020, Ansari et al. 2016, Elangovan et al. 2008).
This chapter aims to review and generalize the observations from different studies on
wastewater remediation of lignocellulosic processing (paper and pulp and textile) industries in
addition to dairy industries from the perspective of the Indian ecosphere. Additionally, the well-
established capabilities of E. crassipes can assist in the remediation of toxic pollutants, including
heavy metals and organic pollutants and other hazardous industrial pollutants, apart from its use in
various documented sustainable practises.
240 Bioremediation for Sustainable Environmental Cleanup
These keywords-oriented text mining studies explored the exploitation of aquatic plants or
macrophytes as a phytoremediation agent for heavy metal contaminated wastewater resources, i.e.,
industrial wastewater. Therefore, interlinking these industrial pollutants and the WH can unearth a
novel sustainable phenomenon, where other aquatic macrophytes can be utilized for such integrated
sustainable practises (heavy metal and other organic pollutants removal or bioremediation).
Phytodegradation
The maximum permissible limit of N, P and K for growth are 20, 30 and 50 ppm, respectively
(Kumar and Chauhan 2019).
WH’s ability to cope with difficult environmental conditions makes it desirable for the
remediation process. During the 1970s and 1980s, WH was generally used for wastewater treatment
as the plant roots are very useful for the absorption of nutrients. The nutrients absorbed by the plant
are stored for growth (Alves et al. 2003, Reddy et al. 1983, Reddy et al. 1987). It can absorb nitrogen
in two stages, first from the medium and another from the sediment. It can also remove P, K and
other organic pollutants, including heavy metals (Chua 1998). WH can control algal blooms and
water quality (Barry 1998). It removes a large number of pollutants, such as BOD, N, COD, P, DO,
heavy metals, etc. (Gupta et al. 2012, Rezania et al. 2015) from the water bodies.
The ability of WH to grow in heavily polluted water makes it favored among plants that use it
in the phytoremediation process (Mahalakshmi et al. 2019, Sarkar et al. 2017, Ebel et al. 2007, Roy
and Hänninen 1994). According to Malik (2007) and Smolyakov (2012) a medium sized WH plant
can proliferate/spread 2 million/hectares. WH grows in water, which is enriched with nutrients, but
fails to grow in saline coastal waters (Jafari 2010, Rezania et al. 2015). Based on dry weight, the
growth rate of WH is 0.04 – 0.08 kg dry weight/m2/d, while the rate based on the surface coverage
is, 1.012 – 1.077 m2d–1 (Gopal 1987, Rezania et al. 2015).
TDS and TSS, phenols and lignins must be minimized through different treatment processes to meet
the pollution standards (Ashrafi et al. 2015).
India’s first paper mill was established at Serampore, West Bengal, in 1812, and hosts the 15th
largest paper and pulp industry in the world. The per capita, paper consumption in India is 5.5 kg y–1
which is 10 times lower than the global average consumption, i.e., 50 kg y–1. The P&P industries in
India are classified into three categories as per the production capacity (tons per annum, TPA), i.e.,
large scale (greater than 33,000 TPA), medium scale (between 10,000 and 33,000 TPA) and small
scale (less than 10,000 TPA). In India mainly, four types of broad categories of paper are produced
and consumed which include firstly, cultural type paper which comes in the form of cream woven,
maplitho, bond paper, chromo paper. Secondly, industrial paper is manufactured in the form of kraft
paper, single layer paper board, multilayer board and duplex board. Thirdly specialty paper which
is security paper, grease proof paper, electrical grades of paper), and fourthly, newsprint (glazed,
non-glazed) (MOEF 2010).
Wastewater generated from the P&P industries contains a high amount of organic matter,
COD, BOD, TSS, Lignin, AOX, Chlorides, PO43–, NO3–, phenols, etc. The study by Sharma et al.
(2021) observed the high amount (693 mg/l) of phenol in the P&P industry effluent apart from a
high amount of PO43– and AOX, whereas Tarlan et al. (2002) observed 267 mg/l and 46.3 mg/l,
respectively of these pollutants. Table 14.1 lists the different pollutants generated in P & P industry
and Table 14.2 shows the Indian standards for Wastewater Generation from P&P industry as given
by CPCB (2000).
Kraft paper, writing paper and hardboard manufacturing industries produce heavy metals like
Fe, Mn, Zn, Cu, Cr, Cd, Ni and Pb, along with organic wastes (Sharma et al. 2020). Different heavy
metals released from different paper-making processes are described in Table 14.3. Heavy metal
tends to bind with lignocellulosic waste and forms complexes like Persistent Organic Pollutants
(POPs), which are a severe threat to the environment and they tend to accumulate in the food chain
(Singh et al. 2015).
244
Type of pH EC (us Lignin Colour COD BOD TDS TSS TS Adsorbabale Chlorides Phosphate Nitrogen Phenol Potassium References
Water cm–1) (mgL–1) (pcu) (mgL–1) (mg (mg (mgL–1) (mgL–1) Organic (mgL–1) (mg L–1) (mg L–1) (mgL–1) (mg L–1)
Table 14.2. Wastewater generation standard for Paper and Pulp industry (CPCB 2000).
industries has different standards as per different pollution control authorities worldwide. As per the
Central Pollution Control Board of India, the standards for different textile wastewater effluents are
described in Table 14.4.
Textile wastewater contains pollutants like inorganic, organic and polymeric products (Brown
and Laboureur 1983). The pollutants include dyes, coloring agents, etc. The dye and coloring agents
are not only harmful to the biological ecosystems, they also block sunlight by covering the surface
layer of soil and water bodies (Choi et al. 2004). Table 14.5 summarizes the pollutants present and
their concentrations in the textile water from different studies conducted worldwide. Wastewaters
from textile industries have shown different levels of various inorganic pollutants, i.e., NH3, PO43–
and N, along with high DO, BOD, COD, TSS, TDS, EC, etc. According to a study by Adelodun
et al. (2021) and Lin and Peng (1994) large amounts of oil and grease are also present in the textile
wastewater.
Along with organic pollutants, textile wastewater is a contributor to heavy metals like Al, Cd,
Cu, Ni, Zn, Pb, Fe, Cr and Hg, etc. A study by Türksoy et al. (2021) has observed a high amount of
Cu (643.74), Zn (398.28) and Ni (95.56). Another study by Sungur and Gülmez (2015) showed the
highest amount of Al and Fe (103.13 mgL–1 and 80.13 mg kg–1) during the wet digestion of cotton
in the textile industry. These authors also found high amounts of Cd, Pb and Cr of 19.43, 189.4 and
382.3 mg kg–1, respectively, in textile sludge. The different concentrations of heavy metals in textile
wastewater observed by different studies are described in Table 14.6.
Cr Cd Mn Ni Pb Fe Cu Zn Hg
Type of Water References
mg kg–1 mg kg–1 mg kg–1 mg kg–1 mg kg–1 mg kg–1 mg kg–1 mg kg–1 mg kg–1
Paper mill sludge 187.3 8.75 _ _ 62.5 _ 144.4 _ _ Suthar et al. 2014
Pulp and paper
Sthiannopkao
industrial wastes 10.11 0.683 184.25 10.39 31.03 199.97 8.748 8.748 0.140
and Sreesai 2009
(Lime mud)
Pulp and paper
Sthiannopkao
industrial wastes 1.210 0.40 42.88 0.403 20.70 37.77 3.360 37.790 0.085
and Sreesai 2009
(Recovery boiler ash)
Pulp and paper mill
Sharma et al.
effluents sludge _ 9.11 ± 0.01 18.27 ± 0.20 5.21 ± 0.04 _ 98.30 ± 1.80 3.21 ± 0.01 51.00 ± 0.00 0.014 ± 0.80
Table 14.4. Wastewater generation standard for Textile industries (mg L–1).
pH DO BOD COD TDS TSS EC (Mohm Ammonia Phosphate Copper Zinc Oil and Chromium References
mg L–1 mg L–1 mg L–1 mg L–1 mg L–1 cm–1) N (Nh3-N)V mg L–1 mg L–1 mg L–1 Grease (mg L–1)
mg L–1 mg L–1
Textile 11.63 0.6 395 1926 4210 3.87 _ _ _ _ _ _ Mahajan et al.
Industry 2019
Wastewater 8.1 _ 48.7 178 _ 5.61 12.14 _ _ _ _ 3.741 Apritama et al.
2020
13.04 _ 1048 2300 4500 _ 3.2*1 _ _ _ _ _ _ Anjana and
Thanga 2011
9.87 3.8 347 2230 1450 370 _ 2.5 3.78 2.45 _ _ _ Ugya 2017
10.7 _ _ 956 _ _ 0.082* 2
5.04 45.7 _ _ 83.3 _ Adelodun et al.
250
Types of Water Al Cd Cu Ni Zn Pb Fe Cr Hg References
Textile effluent _ 0.002 0.078 _ 1.868 0.018 1.122 0.090 3.382 Panigrahi and
Santhoskumar 2020
Textile industry effluent _ 0.016 _ _ 1.02 0.61 1.63 _ _ Odipe et al. 2018
Textile industry effluent _ 0.66 _ _ 1.74 1.02 2.15 _ _ Odipe et al. 2018
Wet digestion of cotton 103.13*1 11.86*1 5.84*1 2.19*1 _ 23.44*1 80.13*1 0.97*1 _ Sungur and Gülmez
2015
Textile Sludge _ 19.43*1 265*1 66.2*1 328*1 189.4*1 _ 382.3*1 0.9*1 Sungur and Gülmez
2015
*1
mg kg–1
Table 14.7. Characterization of dairy industry wastewater.
Type of pH EC TSS TDS TS COD DO BOD SO4 or PO4 or NO3 CO32– Fat Protein Lactose References
Water (µs cm–1) (mg L–1) (mg L–1) (mg L–1) (mg L–1) (mg L–1) (mg L–1) Total S Total P (mg L–1) (mg L–1) (g L–1) (g L–1) (g L–1)
(mg L–1) (mg L–1)
Dairy 4.6 ± 0.2 265 ± 1 420 ± 1 123 ± 05 _ 118 ± 5 1.24 ± 1200 ± 11 40.80 ± 0.37 ± 0.04 0.49 ± 28 ± 0.3 _ _ _ Ondiba et al.
industry 0.4 0.4 0.02 2022)
Dairy 7.1 – _ 221 – _ 800 – 438 – _ _ 17 – 65 17 – 82 _ _ _ _ _ Tait et al. 2021
effluent 8.22 2,996 27,000 5,044
Dairy plant 8.23 ± _ _ _ 0.50 ± 21.50 ± _ 4.50 ± _ _ _ _ _ 0.01 ± 0.33 ± Alalam et al.
cleaning 0.07 0.04*1 0.71 0.71 0.00*1 0.03*1 2021
water
Dairy plant 12.45 ± _ _ _ 3.12 ± 354.50 ± _ _ _ _ _ _ 0.03 ± 0.15 ± 2.22 ± Alalam et al.
252
Industries pH BOD3/BOD5 SS Oil and Grease COD Total Nitrogen Total Phosphorus Total Ammonia Waste Water References
14.7 Conclusion
Water hyacinth has a high potential for removing many organic pollutants and heavy metals from
wastewater under different climatic conditions. Apart from the organic pollutants, N, P, Na, Ca, N,
it helps to reduce many toxic non-essential heavy metals like Al, As, Au, Cd, Cr, Hg, Pb, Pd, Pt,
Sb, Te and Tl from different industrial effluents. Moreover, the study presents characteristics of
the wastewater produced from the different lignocellulosic wastewater-producing industries, i.e.,
paper and pulp mills, dairy and textile industries. However, the dairy industry wastewater has fewer
heavy metals and other pollutants than textile and paper and pulp industries. Different treatment
approaches are taken for removing/minimizing the pollutants. Phytoremediation using water
hyacinth would be one of the cheapest, most cost-effective and environment-friendly approaches to
remediate organic and inorganic pollutants, including heavy metals. Some additional investigation
on uptake efficiency, appropriate climatic conditions and post-harvest treatment processes are
required to develop practical large-scale approaches to phytoremediation using water hyacinth.
254 Bioremediation for Sustainable Environmental Cleanup
pH 7.82 7.29 60 d
Na 285.44 150.33 60 d
K 175.50 96.37 60 d
Ca 435.80 305.80 60 d
Mg 148.35 66.40 60 d
Cd 2.45 1.34 60 d
Cr 1.38 0.69 60 d
Cu 5.64 2.94 60 d
Fe 8.95 4.86 60 d
Mn 3.66 1.42 60 d
Ni 1.74 0.73 60 d
Pb 1.02 0.36 60 d
Zn 6.90 3.10 60 d
*1
mmho cm–1, *2ds m–1
Acknowledgement
Apurba Koley is thankful to the BBSRC, United Kingdom, for granting funding from the BEFWAM
project: Bioenergy, Fertilizer and Clean water from Invasive Aquatic macrophytes [Grant Ref: BB/
S011439/1] for financial support and research fellowship. Sandipan Banerjee and Nitu Gupta thank
the Department of Biotechnology, Govt. of India, for granting DBT Twinning Project and Research
Fellowship [No. BT/PR25738/NER/95/1329/2017 dated December 24, 2018]. While conducting
the study, the authors are thankful for the support from, Dr. Aishiki Banerjee, Ms. Sneha Banerjee.
256 Bioremediation for Sustainable Environmental Cleanup
References
Abedinzadeh, N., M. Shariat, S. M. Monavari and A. Pendashteh. 2018. Evaluation of color and COD removal
by Fenton from biologically (SBR) pre-treated pulp and paper wastewater. Process Saf. Environ. Prot. 116:
82–91.
Adelodun, A. A., O. Temitope, N. O. Afolabi, A. S. Akinwumiju, E. Akinbobola and U. O. Hassan. 2021.
Phytoremediation potentials of Eichhornia crassipes for nutrients and organic pollutants from textile
wastewater. Int. J. Phytoremediat. 23(13): 1333–1341.
Adesra, A., V. K. Srivastava and S. Varjani. 2021. Valorization of dairy wastes: integrative approaches for value added
products. Indian J. Microbiol. 61(3): 270–278.
Ajayi, T. O. and A. O. Ogunbayo. 2012. Achieving environmental sustainability in wastewater treatment by
phytoremediation with water hyacinth (Eichhornia crassipes). J. Sustain. Dev. 5(7): 80.
Alalam, S., F. Ben-Souilah, M. H. Lessard, J. Chamberland, V. Perreault, Y. Pouliot, S. Labrie and A. Doyen. 2021.
Characterization of chemical and bacterial compositions of dairy wastewaters. Dairy. 2(2): 179–190.
Al-Qurainy, F. and A. Abdel-Megeed. 2009. Phytoremediation and detoxification of two organophosphorous
pesticides residues in Riyadh area. World Appl. Sci. J. 6(7): 987–998.
Alves, E., L. R. Cardoso, J. Savroni, L. C. Ferreira, C. S. F. Boaro and A. C. Cataneo. 2003. Avaliações fisiológicas e
bioquímicas de plantas de aguapé (Eichhornia crassipes) cultivadas com níveis excessivos de nutrientes. Planta
Daninha. 21: 27–35.
Anipeddi, M., S. Begum and G. R. Anupoju. 2022. Integrated technologies for the treatment of and resource recovery
from sewage and wastewater using water hyacinth. Biofuel Bioprod. Biorefin, pp. 293–314. Elsevier,
Anjana, S. and V. S. G. Thanga. 2011. Phytoremediation of synthetic textile dyes. Asian J. Microbiol. Biotechnol.
Environ. Sci. 13: 31–34.
Ansari, A. A., R. Gill, L. Newman, S. S. Gill and G. R. Lanza. 2016. Phytoremediation: Management of Environmental
Contaminants. 6: 1–576. https://doi.org/10.1007/978-3-319-99651-6.
Apritama, M. R., I. Suryawan, A. S. Afifah and I. Y. Septiariva. 2020. Phytoremediation of effluent textile wwtp for
NH3-N and Cu reduction using pistia stratiotes. Plant Archives. 20(1): 2384–2388.
Arivoli, A., R. Mohanraj and R. Seenivasan. 2015. Application of vertical flow constructed wetland in treatment of
heavy metals from pulp and paper industry wastewater. Environ. Sci. Pollut. Res. 22(17): 13336–13343.
Ashrafi, O., L. Yerushalmi and F. Haghighat. 2015. Wastewater treatment in the pulp-and-paper industry: a review of
treatment processes and the associated greenhouse gas emission. J. Environ. Manag. 158: 146–157.
Assche, V. F. and H. Clijsters. 1990. Effects of metals on enzyme activity in plants. Plant Cell Environ. 13(3): 195–
206.
Azadi Aghdam, M., H. R. Kariminia and S. Safari. 2016. Removal of lignin, COD, and color from pulp and paper
wastewater using electrocoagulation. Desalin. Water Treat. 57(21): 9698–9704.
Babu, B. R., A. K. Parande, S. Raghu and T. P. Kumar. 2007. Cotton textile processing: waste generation and effluent
treatment. J. Cotton Sci. 11: 141–153.
Barry, A. C. P. 1998. Preliminary investigation of an integrated aquaculture–wetland ecosystem using tertiary-treated
municipal wastewater in Los Angeles County, California. Ecol. Eng. 10(4): 341–354.
Basha, S. A. and K. Rajaganesh. 2014. Microbial bioremediation of heavy metals from textile industry dye effluents
using isolated bacterial strains. Int. J. Curr. Microbiol. Appl. Sci. 3: 785–794.
Basu, A., A. K. Hazra, S. Chaudhury, A. B. Ross and S. Balachandran. 2021. State of the art research on sustainable
use of water hyacinth: a bibliometric and text mining analysis. In Informatics. Multidisciplinary Digital
Publishing Institute. 8(2): 38.
Bazrafshan, E., H. Moein, F. KordMostafapour and S. Nakhaie. 2013. Application of electrocoagulation process for
dairy wastewater treatment. J. Chem. 7–10.
Bhavsar, S. R., V. R. Pujari and V. V. Diwan. 2010. Potential of phytoremediation for dairy wastewater treatment. J.
Civ. Eng. Manag. 16–23.
Bhavsar, S. R., V. R. Pujari and V. V. Diwan. 2012. Potential of phytoremediation for dairy wastewater treatment.
IIOSR J. Mech. Civ. Eng. 16–23.
Bortoluzzi, A. C., J. A. Faitão, M. Di Luccio, R. M. Dallago, J. Steffens, G. L. Zabot and M. V. Tres. 2017. Dairy
wastewater treatment using integrated membrane systems. J. Environ. Chem. Eng. 5(5): 4819–4827.
Brião, V. B., A. C. Vieira Salla, T. Miorando, M. Hemkemeier and D. P. CadoreFavaretto. 2019. Water recovery
from dairy rinse water by reverse osmosis: giving value to water and milk solids. Resour. Conserv. Recycl.
140: 313–323.
Brown, D. and P. Laboureur. 1983. The aerobic biodegradability of primary aromatic amines. Chemosphere. 12(3):
405–414.
Phytoremediation of Industrial Wastewater using Water Hyacinth 257
Cai, C., H. Wang and Z. Zhang. 2004. Removal of Cu, Pb, Cd, Zn and Fe by water hyacinth. J. Leshan Teachers
College. 19(6): 69–72.
Chandel, A. K., V. K. Garlapati, A. K. Singh, F. A. F. Antunes and S. S. da Silva. 2018. The path forward for
lignocellulose biorefineries: bottlenecks, solutions, and perspective on commercialization. Bioresour. Technol.
264: 370–381. doi: 10.1016/j.biortech.2018.06.00.
Chandra, R., S. Yadav and S. Yadav. 2017. Phytoextraction potential of heavy metals by native wetland plants growing
on chlorolignin containing sludge of pulp and paper industry. Ecol. Eng. 98: 134–145.
Choi, J. W., H. K. Song, W. Lee, K. K. Koo, C. Han and B. K. Na. 2004. Reduction of COD and color of acid and
reactive dyestuff wastewater using ozone. Korean J. Chem. Eng. 21(2): 398–403.
Chua, H. 1998. Bio-accumulation of environmental residues of rare earth elements in aquatic flora Eichhornia
crassipes (Mart.) Solms in Guangdong Province of China. Sci. Total Environ. 214(1-3): 79–85.
Chuphal, Y., V. Kumar and I. S. Thakur. 2005. Biodegradation and decolorization of pulp and paper mill effluent by
anaerobic and aerobic microorganisms in a sequential bioreactor. World J. Microbiol. Biotechnol. 21(8-9):
1439–1445.
CPCB. 2000. Environmental Standards for Ambient Air, Automobiles, Fuels, Industries and Noise, Central Pollution
Control Board. Ministry of Environment and Forest.
Dhall, V. P. 2014. Biological approach for the treatment of pulp and paper industry effluent in sequence batch reactor.
J. Bioremediat. Biodegrad. 5(3).
Dhir, B. 2013. Phytoremediation: role of aquatic plants in environmental clean-up. 14: 1–111. Springer, New Delhi.
Djingova, R. and I. v Kuleff. 2000. Instrumental techniques for trace analysis. In Trace metals in the Environment.
4: 137–185. Elsevier.
Doble, M. and A. Kumar. 2005. Biotreatment of industrial effluents.1st ed. Burlington: Butterworth-Heinemann.
Elsevier.
Ebel, M., M. W. Evangelou and A. Schaeffer. 2007. Cyanide phytoremediation by water hyacinths (Eichhornia
crassipes). Chemosphere. 66(5): 816–823.
Ekanayake, M. S., D. Udayanga, I. Wijesekara and P. Manage. 2021. Phytoremediation of synthetic textile dyes:
biosorption and enzymatic degradation involved in efficient dye decolorization by Eichhornia crassipes
(Mart) Solms and Pistia stratiotes L. Environ. Sci. Pollut. Res. 16: 20476–20486.
Elangovan, R., L. Philip and K. Chandraraj. 2008. Biosorption of chromium species by aquatic weeds: kinetics and
mechanism studies. J. Hazard. Mater. 152(1): 100–112.
Ensley, B. D. 2000. Rationale for the Use of Phytoremediation. Phytoremediation of toxic metals: Using plants to
Clean-up the Environment. John Wiley Publishers: New York.
FAO. 2020. The State of World Fisheries and Aquaculture 2020. Sustainability in Action.
Fletcher, J., N. Willby, D. M. Oliver and R. S. Quilliam. 2020. Phytoremediation using aquatic plants. In
Phytoremediation, pp. 205–260. Springer, Cham.
Gamage, N. S. and P. A. J. Yapa. 2001. Use of water Hyacinth (Eichhornia crassipes (Mart) Solms) in treatment
systems for textile mill effluents-a case study. J. Natl. Sci. Found. Sri Lanka. 29: 1–2.
Gopal, B. 1987. Water hyacinth. Elsevier Science Publishers.
Gupta, A. and R. Gupta. 2019. Treatment and recycling of wastewater from pulp and paper mill. In Advances in
Biological Treatment of Industrial Waste Water and their Recycling for a Sustainable Future. pp. 13–49.
Springer, Singapore.
Gupta, P., S. Roy and A. B. Mahindrakar. 2012. Treatment of water using water hyacinth, water lettuce and vetiver
grass–a review. System. 49: 50.
Hao, O. J., H. Kim and P. C. Chiang. 2000. Decolorization of wastewater. Crit. Rev. Environ. Sci. Technol. 30(4):
449–505.
Holik, H. (Ed.). 2006. Handbook of paper and board. John Wiley and Sons.
Holkar, C. R., A. J. Jadhav, D. V. Pinjari, N. M. Mahamuni and A. B. Pandit. 2016. A critical review on textile
wastewater treatments: possible approaches. J. Environ. Manag. 182: 351–366.
Hooda, R., N. K. Bhardwaj and P. Singh. 2015. Screening and identification of ligninolytic bacteria for the treatment
of Pulp and Paper Mill effluent. Water Air Soil Pollut. 226(9).
Huang, B. and H. Xu. 2008. Water hyacinth on ecological damage and phytoremediation. Guangdong Water Resour.
Hydropower. (3): 1–3, 11.
Jadia, C. D. and M. H. Fulekar. 2009. Phytoremediation of heavy metals: recent techniques. Afr. J. Biotechnol. 8(6).
Jafari, N. 2010. Ecological and socio-economic utilization of water hyacinth (Eichhornia crassipes Mart Solms). J.
Appl. Sci. Environ. Manag. 14(2).
Kamal, M., A. E. Ghaly, N. Mahmoud and R. CoteCôté. 2004. Phytoaccumulation of heavy metals by aquatic plants.
Environ. Int. 29(8): 1029–1039.
258 Bioremediation for Sustainable Environmental Cleanup
Kaur, V., G. Sharma and C. Kirpalani. 2018. Agro-potentiality of dairy industry effluent on the characteristics of
Oryza sativa L. (Paddy). Environ. Technol. Innov. 12: 132–147.
Khansorthong, S. and M. Hunsom. 2009. Remediation of wastewater from pulp and paper mill industry by the
electrochemical technique. Chem. Eng. J. 151(1–3): 228–234.
Kolawole, B. S. 2001. Cleaning of effluent from textile Industry by water hyacinth (B.Sc thesis). University of
Agriculture, Abeokuta, Ogun state.
Koley, A., D. Bray, S. Banerjee, S. Sarhar, R. Ghosh Thahur, A. K. Hazra, N. C. Mandal et al. 2022. Water hyacinth
(Eichhornia crassipes) a sustainable strategy for heavy metals removal from contaminated waterbodies.
pp. 95–114. In: A. Malik, M. K. Kidwai and V. K. Garg [Eds.]. Bioremediation of Toxic Metal(loid)s. CRC
Press, London
Kumar, P. and M. S. Chauhan. 2019. Adsorption of chromium (VI) from the synthetic aqueous solution using
chemically modified dried water hyacinth roots. J. Environ. Chem. Eng. 7(4): 103218.
Kumar, R., K. S. Rajeev and P. S. Anirudh. 2017. Cellulose based grafted biosorbents-Journey from lignocellulose
biomass to toxic metal ions sorption applications—a review. J. Mol. Liq. 232: 62–93.
Kumar, S., T. Saha and S. Sharma. 2015. Treatment of pulp and paper mill effluents using novel biodegradable
polymeric flocculants based on anionic polysaccharides: a new way to treat the waste water. Int. Res. J. Eng.
Technol. 2(4): 1–14.
Kumar, V., A. K. Chopra, J. Singh, R. K. Thakur, S. Srivastava and R. K. Chauhan. 2016. Comparative assessment
of phytoremediation feasibility of water caltrop (Trapa natans L.) and water hyacinth (Eichhornia crassipes
Solms) using pulp and paper mill effluent. Arch. Agri. Sci. J. 1(1): 13–21.
Kumar, V., J. Singh and P. Kumar. 2020. Regression models for removal of heavy metals by water hyacinth (Eichhornia
crassipes) from wastewater of pulp and paper processing industry. Environ. Sustain. 3(1): 35–44.
Kushwaha, J. P., V. C. Srivastava and I. D. Mall. 2011. An overview of various technologies for the treatment of dairy
wastewaters. Crit. Rev. Food Sci. Nutr. 51(5): 442–452.
Lawrence, W., H. Yung-Tse and S. Nazih. 2005. Physicochemical treatment processes. pp. 141–154. In: L. K. Wang,
Y. T. Hung and N. K. Shammas [Eds.]. Handbook of Environmental Engineering. Humana Press.
Lin, S. H. and C. F. Peng. 1994. Treatment of textile wastewater by electrochemical method. Water Res. 28(2):
277–282.
Liu, J., F. Lin, Y. Wang, Z. Xu and X. Zhang. 2003. Absorption processes of macrophyte root on polycyclic aromatic
hydrocarbons (naphthalene). Environ. Sci. Technol. 26(1): 32–34.
Liu, R. R., Q. Tian, B. Yang and J. H. Chen. 2010. Hybrid anaerobic baffled reactor for treatment of desizing
wastewater. Int. J. Environ. Sci. Technol. 7(1): 111–118.
Lu, X., L. Liu, R. Fan, J. Luo, S. Yan, Z. Rengel and Z. Zhang. 2017. Dynamics of copper and tetracyclines during
composting of water hyacinth biomass amended with peat or pig manure. Environ. Sci. Pollut. Res. 24(30):
23584–23597.
Lu, X., Y. Gao, J. Luo, S. Yan, Z. Rengel and Z. Zhang. 2014. Interaction of veterinary antibiotic tetracyclines and
copper on their fates in water and water hyacinth (Eichhornia crassipes). J. Hazard. Mater. 280: 389–398.
Mahajan, P., J. Kaushal, A. Upmanyu and J. Bhatti. 2019. Assessment of phytoremediation potential of Chara vulgaris
to treat toxic pollutants of textile effluent. J. Toxicol. ID 8351272 | https://doi.org/10.1155/2019/8351272.
Mahalakshmi, R., C. Sivapragasam and S. Vanitha. 2019. Comparison of BOD 5 Removal in water hyacinth and
duckweed by genetic programming. Int. J. Inf. Commun. Technol. 401–408.
Mahesh, S., B. Prasad, I. D. Mall and I. M. Mishra. 2006. Electrochemical degradation of pulp and paper mill
wastewater. Part 1. COD and color removal. Ind. Eng. Chem. Res. 45(8): 2830–2839.
Mahmood, Q., P. Zheng, E. Islam, Y. Hayat, M. J. Hassan, G. Jilani and R. C. Jin. 2005. Lab scale studies on water
hyacinth (Eichhornia crassipes Marts Solms) for biotreatment of textile wastewater. Casp. J. Environ. Sci.
3(2): 83–88.
Malik, A. 2007. Environmental challenge vis a vis opportunity: the case of water hyacinth. Environ. Int. 33(1): 122–
138.
Margoshes, M. and B. L. Vallee. 1957. A cadmium protein from equine kidney cortex. J. Am. Chem. Soc. 79(17):
4813–4814.
Menon, V. and M. Rao. 2012. Trends in bioconversion of lignocellulose: biofuels, platform chemicals and biorefinery
concept. Prog. Energy Combust. Sci. 38: 522–550. doi: 10.1016/j.pecs.2012.02.002.
Ministry of Textiles, Government of India 21–22 Annual Report Ministry of Textiles.
MOEF. 2010. The Ministry of Environment and Forests, Government of India, Technical EIA Guidance Manual for
Pulp and Paper Industries, Prepared by IL and FS Ecosmart Limited, Hyderabad.
Mokhtar, H., N. Morad and F. F. A. Fizri. 2011. Hyperaccumulation of copper by two species of aquatic plants. Int.
Conf. Eng. Environ. Sci. 8: 115–118.
Phytoremediation of Industrial Wastewater using Water Hyacinth 259
Munavalli, G. R. and P. S. Saler. 2009. Treatment of dairy wastewater by water hyacinth. Water Sci. Technol. 59(4):
713–722.
Nahar, K., M. Chowdhury, A. Khair, M. Chowdhury, A. Hossain, A. Rahman and K. M. Mohiuddin. 2018. Heavy
metals in handloom-dyeing effluents and their biosorption by agricultural byproducts. Environ. Sci. Pollut.
Res. 25(8): 7954–7967. https://doi. org/10.1007/s11356-017-1166-9.
Nakamura, Y., T. Sawada, F. Kobayashi and M. Godliving. 1997. Microbial treatment of kraft pulp wastewater
pretreated with ozone. Water Sci. Technol. 35(2-3): 277–282.
Nawirska, A. 2005. Binding of heavy metals to pomace fibers. Food Chem. 90(3): 395–400.
Nemerow, N. L. 2007. Industrial waste treatment. Elsevier science and technology. Burlington, MA: Butterworth-
Heinemann.
Nibret, G., S. Ahmad, D. G. Rao, I. Ahmad, M. A. M. U. Shaikh and Z. U. Rehman. 2019. Removal of methylene blue
dye from textile wastewater using water hyacinth activated carbon as adsorbent: synthesis, characterization and
kinetic studies. In Proceedings of International Conference on Sustainable Computing in Science, Technology
and Management (SUSCOM), Amity University Rajasthan, Jaipur-India.
Nye, P. H. and P. B. Tinker. 1977. Solute movement in the soil-root system. Univ. of California Press.
Odipe, O. E., M. O. Raimi and F. Suleiman. 2018. Assessment of heavy metals in effluent water discharges from
textile industry and river water at close proximity: A comparison of two textile industries from Funtua and
Zaria, North Western Nigeria. Madridge J. Agric. Environ. Sci. 1(1): 1–6.
Ondiba, J. O., C. L. Kanali, B. B. Gathitu and S. N. Ondimu. 2022. Characterisation and quantification of bioprocessing
effluents from coffee, dairy and tanneryplants. Int. J. Agric. Environ. Res. 08(02): 265–277.
Onet, C. 2010. Characteristics of the untreated wastewater produced by food industry. Analele Universităţii Din
Oradea, Fascicula: Protecţia Mediului, XV: 709–714.
Panigrahi, T. and A. U. Santhoskumar. 2020. Adsorption process for reducing heavy metals in Textile Industrial
Effluent with low cost adsorbents. Prog. Chem. Biochem. Res. 3(2): 135–139.
Paraíba, L. C. 2007. Pesticide bioconcentration modelling for fruit trees. Chemosphere. 66(8): 1468–1475.
Pascale, N. C., J. J. Chastinet, D. M. Bila, G. L. Sant’Anna, S. L. Quitério and S. M. R. Vendramel. 2019. Enzymatic
hydrolysis of floatable fatty wastes from dairy and meat food-processing industries and further anaerobic
digestion. Water Sci. Technol. 79(5): 985–992.
Pizzichini, M., C. Russo and C. Di Meo. 2005. Purification of pulp and paper wastewater, with membrane technology,
for water reuse in a closed loop. Desalination 178(1-3): 351–359.
Porwal, R. K. 2015. Cost control opportunities in dairy industry. Indian Dairyman. 67(2): 82–87.
Pramanik, K., S. Mandal, S. Banerjee, A. Ghosh, T. K. Maiti and N. C. Mandal. 2021. Unraveling the heavy metal
resistance and biocontrol potential of Pseudomonas sp. K32 strain facilitating rice seedling growth under Cd
stress. Chemosphere. 274: 129819.
Prasongsuk, S., P. Lotrakul, T. Imai and H. Punnapayak. 2009. Decolourization of pulp mill wastewater using
thermotolerant white rot fungi. Sci. Asia. 35: 37–41.
Qin, H., M. Diao, Z. Zhang, P. M. Visser, Y. Zhang, Y. Wang and S. Yan. 2020. Responses of phytoremediation in
urban wastewater with water hyacinths to extreme precipitation. J. Environ. Manag. 271: 110948.
Rai, P. K. and B. D. Tripathi. 2007. Heavy metals removal using nuisance blue green alga Microcystis in continuous
culture experiment. Environ. Sci. 4(1): 53–59. doi:10.1080/15693430601164956.
Rai, P. K. and B. D. Tripathi. 2008. Heavy metals in industrial wastewater, soil and vegetables in Lohta village,
India. Toxicol. Environ. Chem. 90(2): 247–257.doi:10.1080/02772240701458584.
Rai, P. K. 2009. Heavy metals in water, sediments and wetland plants in an aquatic ecosystem of tropical industrial
region, India. Environ. Monit. Assess. 158(1): 433–457.
Rai, U. N., R. D. Tripathi, P. Vajpayee, V. Jha and M. B. Ali. 2002. Bioaccumulation of toxic metals (Cr, Cd, Pb and
Cu) by seeds of Euryale ferox Salisb (Makhana). Chemosphere. 46(2): 267–272.
Reddy, G. B. and K. R. Reddy. 1987. Nitrogen transformations in ponds receiving polluted water from nonpoint
sources. Am. Soc. Agron. Crop Sci. Soc. Am., and Soil Sci. Soc. Am. 16:(1):1–5.
Reddy, K. R. and J. C. Tucker. 1983. Productivity and nutrient uptake of water hyacinth, Eichhornia crassipes I.
Effect of nitrogen source. Econ. Bot. 37(2): 237–247.
Rezania, S., M. Ponraj, A. Talaiekhozani, S. E. Mohamad, M. F. M. Din et al. 2015. Perspectives of
phytoremediation using water hyacinth for removal of heavy metals, organic and inorganic pollutants in
wastewater. J. Environ. Manag. 163: 125–133.
Rodrigues, A. C., M. Boroski, N. S. Shimada, J. C. Garcia, J. Nozaki and N. Hioka. 2008. Treatment of paper pulp and
paper mill wastewater by coagulation-flocculation followed by heterogeneous photocatalysis. J. Photochem.
Photobiol. A: Chem. 194(1): 1–10.
260 Bioremediation for Sustainable Environmental Cleanup
Roufou, S., S. Griffin, L. Katsini, M. Polańska, J. F. V. Impe and V. P. Valdramidis. 2021. The (potential) impact
of seasonality and climate change on the physicochemical and microbial properties of dairy waste and its
management. Trends Food Sci. Technol. 116: 1–10.
Roy, R., A. N. M. Fakhruddin, R. Khatun and M. S. Islam. 2010. Reduction of COD and pH of textile industrial
effluents by aquatic macrophytes and algae. J. Bangl. Acad. Sci. 34 (1): 9–14.
Roy, S. and O. Hänninen. 1994. Pentachlorophenol: uptake/elimination kinetics and metabolism in an aquatic plant,
Eichhornia crassipes. Environ. Toxicol. Chem. 13(5): 763–773.
Saadia, A. and A. Ashfaq. 2010. Environmental management in pulp and paper industry. J. Industrial Control
Pollution. 26(1).
Safauldeen, S. H., H. A. Hasan and S. R. S. Abdullah. 2019. Phytoremediation efficiency of water hyacinth for batik
textile effluent treatment. J. Ecol. Eng. 20(9): 177–187.
Salt, D. E., M. Blaylock, N. P. Kumar, V. Dushenkov, B. D. Ensley, I. Chet and I. Raskin. 1995. Phytoremediation: a
novel strategy for the removal of toxic metals from the environment using plants. Biotechnol. 13(5): 468–474.
Samecka-Cymerman, A. and A. J. Kempers. 2007. Heavy metals in aquatic macrophytes from two small rivers
polluted by urban, agricultural and textile industry sewages SW Poland. Arch. Environ. Contam. Toxicol.
53(2): 198–206.
Sarkar, M., A. K. M. L. Rahman and N. C. Bhoumik. 2017. Remediation of chromium and copper on water hyacinth
(E. crassipes) shoot powder. Water Resour. Ind. 17: 1–6.
Schnoor, J. L., L. A. Light, S. C. McCutcheon, N. L. Wolfe and L. H. Carreia. 1995. Phytoremediation of organic and
nutrient contaminants. Environ. Sci. Technol. 29(7): 318A–323A.
Sevimli, M. F. 2005. Post-treatment of pulp and paper industry wastewater by advanced oxidation processes. Ozone:
Sci. Eng. 27(1): 37–43.
Sharma, P., S. Tripathi and R. Chandra. 2020. Phytoremediation potential of heavy metal accumulator plants for waste
management in the pulp and paper industry. Heliyon. 6(7): e04559.
Sharma, P., S. Tripathi and R. Chandra. 2021. Highly efficient phytoremediation potential of metal and metalloids from
the pulp paper industry waste employing Eclipta alba (L.) and Alternanthera philoxeroide (L.): Biosorption
and pollution reduction. Bioresour. Technol. 319: 124147.
Show, B. K., S. Balachandran, A. Banerjee, R. GhoshThakur, A. K. Hazra, N. C. Mandal, A. B. Ross, B. Srinivasan
and S. Chaudhury. 2022. Insect gut bacteria: a promising tool for enhanced biogas production. Rev. Environ.
Sci. Biotechnol. 21: 1–25.
Singh, M., M. Pant, S. Diwan and V. Snasel. 2022. Genetic algorithm-enhanced rank aggregation model to measure
the performance of Pulp and Paper Industries. Comput. Ind. Eng. 172: 108548.
Singh, R., S. Singh, P. Parihar, V. P. Singh and S. M. Prasad. 2015. Arsenic contamination, consequences and
remediation techniques: a review. Ecotoxicol. Environ. Saf. 112: 247–270.
Sinha, D., S. Banerjee, S. Mandal, A. Basu, A. Banerjee, S. Balachandran and S. Chaudhury. 2021. Enhanced biogas
production from Lantana camara via bioaugmentation of cellulolytic bacteria. Bioresour. Technol. 340:
125652.
Smolyakov, B. S. 2012. Uptake of Zn, Cu, Pb, and Cd by water hyacinth in the initial stage of water system
remediation. Appl. Geochem. 27(6): 1214–1219.
Sthiannopkao, S. and S. Sreesai. 2009. Utilization of pulp and paper industrial wastes to remove heavy metals from
metal finishing wastewater. J. Environ. Manag. 90(11): 3283–3289.
Sungur, Ş. and F. Gülmez. 2015. Determination of metal contents of various fibers used in textile industry by MP-
AES. J. Spectrosc.
Suthar, S., P. Sajwan and K. Kumar. 2014. Vermiremediation of heavy metals in wastewater sludge from paper and
pulp industry using earthworm Eisenia fetida. Ecotoxicol. Environ. Saf. 109: 177–184.
Tait, S., P. W. Harris and B. K. McCabe. 2021. Biogas recovery by anaerobic digestion of Australian agro-industry
waste: A review. J. Clean. Prod. 299: 126876.
Tarlan, E., F. B. Dilek and U. Yetis. 2002. Effectiveness of algae in the treatment of a wood-based pulp and paper
industry wastewater. Bioresour. Technol. 84(1): 1–5.
The Environment (Protection) Rules. 1986. Ministry of Environment and Forests, Department of Environment, Forest
and Wildlife, New Delhi.
The World Bank Group. 1999. Pollution prevention and abatement handbook, 1998: toward cleaner. The International
Bank for Reconstruction and Development (Washington D.C., United states).
Toczyłowska-Mamińska, R. 2017. Limits and perspectives of pulp and paper industry wastewater treatment–A
review. Renew. Sustain. Energy Rev. 78: 764–772.
Trivedy, R. K. and S. M. Pattanshetty. 2002. Treatment of dairy waste by using water hyacinth. Water Sci. Technol.
45(12): 329–334.
Phytoremediation of Industrial Wastewater using Water Hyacinth 261
Türksoy, R., G. Terzioğlu, İ. E. Yalçin, Ö. Türksoy and G. Demir. 2021. Removal of heavy metals from textile
industry wastewater. Front. Life Sci. RT. 2(2): 44–50.
U.S. Environmental Protection Agency (EPA). 2008. EPA’s 2008 Report on the Environment. National Center for
Environmental Assessment, Washington, DC; EPA/600/R-07/045F.
Uddin, F. 2021. Environmental hazard in textile dyeing wastewater from local textile industry. Cellulose. 28(17):
10715–10739.
Ugya, A. Y., T. S. Imam and A. S. Hassan. 2017. Phytoremediation of textile waste water using Azolla pinnata; a case
study. World J. Pharm. Res. 6(2).
United States Protection Agency (USPA). 2000. Introduction to Phytoremediation. EPA 600/R-99/107. U.S.
Environmental Protection Agency, Office of Research and Development, Cincinnati, OH.
Verma, A. and A. Singh. 2017. Physico-chemical analysis of dairy industrial effluent. Int. J. Curr. Microbiol. Appl.
Sci. 6: 1769–1775.
Vinodhini, M. and C. Soundhari. 2019. Phycoremediation of dairy effluent by using microalgal consortium.
J. Pharm. Innov. 8(9): 128–133.
Voudrias, E. A. and K. S. Assaf. 1996. Theoretical evaluation of dissolution and biochemical reduction of TNT for
phytoremediation of contaminated sediments. J. Contam. Hydrol. 23(3): 245–261.
Wang, B., L. Gu and H. Ma. 2007. Electrochemical oxidation of pulp and paper making wastewater assisted by
transition metal modified kaolin. J. Hazard. Mater. 143(1-2): 198–205.
Wild, E., J. Dent, G. O. Thomas and K. C. Jones. 2005. Direct observation of organic contaminant uptake, storage,
and metabolism within plant roots. Environ. Sci. Technol. 39(10): 3695–3702.
Xia, H. L., L. H. Wu and Q. N. Tao. 2002. Phytoremediation of some pesticides by water hyacinth (Eichhornia
crassipes Solms). J. Zhejiang Univ. - Agric. Life Sci. 28(2): 165–168.
Xia, H., L. Wu and Q. Tao. 2003. A review on phytoremediation of organic contaminants. Ying yong sheng tai xue
bao. J. Appl. Ecol. 14(3): 457–460.
Yan, S. H. and J. Y. Guo. 2017. Water hyacinth: Environmental challenges, management and utilization. In Water
Hyacinth: Environmental Challenges, Management and Utilization. CRC Press, 1–327.
Yaseen, D. A. and M. Scholz. 2019. Textile dye wastewater characteristics and constituents of synthetic effluents: a
critical review. Int. J. Environ. Sci. Technol. 16(2): 1193–1226.
Zavoda, J., T. Cutright, J. Szpak and E. Fallon. 2001. Uptake, selectivity, and inhibition of hydroponic treatment of
contaminants. J. Environ. Eng. 127(6): 502–508.
Zhang, Z., Z. Rengel, K. Meney, L. Pantelic and R. Tomanovic. 2011. Polynuclear aromatic hydrocarbons (PAHs)
mediate cadmium toxicity to an emergent wetland species. J. Hazard. Mater. 189(1-2): 119–126.
Zhao, X., L. Zhang and D. Liu. 2012. Biomass recalcitrance. Part II: Fundamentals of different pre‐treatments to
increase the enzymatic digestibility of lignocellulose. Biofuel Bioprod. Biorefin. 6(5): 561–579.
Zhou, W., L. Tan, D. Liu, H. Yan, M. Zhao and D. Zhu. 2005. Research advances of Eichhornia crassipes and it’s
utilization. J. Anhui Agric. Univ. 24(4): 423–428.
Chapter 15
Biohybrids for Environmental
Remediation and Biosensing
Concept, Synthesis and Future
Prospective
Archana Mishra,1,2,* Ayushi Rastogi 3 and
Avanish Singh Parmar 4
15.1 Introduction
There is constant release of a wide range of pollutants (heavy metals, pesticides, etc.,) into water
bodies which leads to pollution of water resources. Water pollution is a serious concern as it causes
scarcity of drinking water worldwide. Organic as well as metal pollutants present in water bodies
are harmful to all living systems and environment. Several processes such as adsorption (Douglas
et al. 2016), biodegradation (Li et al. 2016), coagulation (Zhu et al. 2016) and photocatalytic oxidation
(Chong et al. 2010) have been used for the removal of pollutants and efficient wastewater treatment,
however; these have their own limitations, like generation of waste products, poor removal capacity,
high energy demand and high cost.
Microorganisms are well known for wastewater treatment as these degrade a variety of
substrates for their consumption using their metabolic diversity. A wide range of microorganism
like Aspergillus niger (Vassilev et al. 1997), Pseudomonas aeruginosa (Shukla et al. 2014) and
Rhodopseudomonas sphaeroides (Liu et al. 2015) has been used for wastewater treatment. However,
limitations like slow biodegradation processes, difficulty in recovering cells, sensitivity towards the
surrounding environment and poor activities of recovered cells persist. There is an urgent need
to find economical cost-effective solutions for remediation of these pollutants. The combination
of the microorganism with suitable support could be a low-cost, environmentally benign and
efficient technique (Mishra et al. 2014, Oh et al. 2016) to achieve the desired outcomes is needed.
1
Nuclear Agriculture and Biotechnology Division, Bhabha Atomic Research Centre, Trombay, Mumbai-400 085, India.
2
Homi Bhabha National Institute, Anushakti Nagar, Mumbai-400 094, India.
3
Department of Humanities and Applied Sciences, School of Management Science (SMS) Institute of Technology, College
of Engineering, Lucknow – 226001, Uttar Pradesh, India.
4
Department of Physics, Indian Institute of Technology (BHU) Varanasi,Uttar Pradesh, India.
* Corresponding author: [email protected], [email protected]
Biohybrids for Environmental Remediation and Biosensing 263
15.2 Biohybrids
In the field of material science, rapid development has been made in the past few decades wherein
biological sciences have contributed significantly. Biohybrid materials have played an important
role. A biohybrid material is composed of two components. One is the biologically active component
which includes microorganisms, living cells, enzymes, etc., and the other supports (organic/
inorganic nanomaterials) (Figure 15.1). Biomolecules provide a process such as their functions
of synthesis, sensing, secretion, etc. However, support materials enable protection and stability to
the biomolecules (Ouyang et al. 2020). As advantages of biomolecules and support materials are
combined, biohybrids exhibit improved characteristics over conventional materials and offer dual
functionality.
For the synthesis of biohybrid materials, many biomolecules (Mishra et al. 2017, Mishra
et al. 2020a, Shukla et al. 2020, Ouyang et al. 2020) and as support material various polymers,
nanoparticles, etc., have been used. A support material is present as a coating over the surface of
biomolecules and gives protection against a harsh microenvironment. In order to develop biohybrids
different techniques like sol-gel technique, moulding, electrospinning, spray drying, microfluidics,
3D printing have been widely applied. Using these methods, biohybrids with various morphologies
Figure 15.1. Biohybrid materials, its components and potential field of applications.
264 Bioremediation for Sustainable Environmental Cleanup
like microparticles, fibres, sheets and scaffolds can be synthesized. Biomolecules present in
biohybrid materials could perform required functions and enhance the applicability of biohybrids
for drug delivery, biomedical engineering, biosensing, wearable devices, etc. (Mishra et al. 2017,
Mishra et al. 2021a, Rivera-Tarazona et al. 2021, Wang et al. 2022). Several reviews are available
on biohybrid materials (Nguyen et al. 2018, Mishra et al. 2020b, Wang et al. 2022).
and
The association of living cells with support will certainly improve the application of both
the components. Living cells offer functionality, however they lack stability. On the other hand, a
support has high mechanical and chemical stability, while it is deficient in functionality. A biohybrid
wherein both (biomolecule and support) are present, exhibit a dual and improved property. The
association of living cells with a support can be achieved using various methods such as cells can be
used as a template and support can be assembled on it or a biomolecule can be entrapped in support.
The method was selected in such a way that it would impart a positive effect on the applicability
of biomolecule (Zhu et al. 2019). A biosensor was developed using Sphingomonas sp. cells. The
microbial cells have periplasmic enzyme which hydrolyzes methyl parathion pesticide, however
the poor storage stability of enzyme was of concern. Thus, functionalized silica nanoparticles were
assembled on the surface of cells and it was observed that the storage stability improved significantly
(Mishra et al. 2017). In order to address stress resistance of living cells, inorganic components with
tuneable properties and high stability are combined , which provides protection against stresses.
A biocompatible and porous support is preferred which helps in transport of nutrients and required
gases (Gerber et al. 2012). The properties of biocomponents in a biohybrid are regulated by the
properties of both the biomolecule and the support. The interactions between both the components
govern the biochemical and kinetic properties of biomolecule. Support materials could be classified
into two major groups, i.e., organic support and inorganic support.
polymer with the required characteristics and desired functional groups is a costly and time-
consuming method which also generates chemical wastes.
15.4.1.2 Biopolymers
In comparison to synthetic polymers, biopolymers are naturally available biocompatible supports.
Many biopolymers like alginate, chitosan, cellulose, and pectin, etc., are commercially available
and used as supports. Biopolymers possess characters like biocompatibility, availability of various
functional groups, biodegradability to harmless products, non-toxicity and affinity towards proteins.
These characteristics increase the applicability of biopolymers as support for biomolecules and
provide a favourable microenvironment to the attached biomolecule which helps to maintain/
improve biocatalytic aspect of biomolecule.
15.4.1.2a. Alginate: Alginate is a biopolymer which is widely applied for developing supports
of various morphologies and immobilization of biomolecules. Earlier lipase was immobilized in
calcium alginate beads and showed high immobilization yields (Betigeri and Neau 2002), however
reusability was poor due to the leaching of the enzyme from the support. The application of alginate
is limited due to its sensitivity towards divalent ions and low mechanical stability of alginate gels
and substrate diffusional limitations (Coradin et al. 2003). To address the issue, a support comprising
nanosilica-sodium alginate was prepared and chemotherapeutic drug doxorubicin was entrapped
(Mishra et al. 2021a). It was observed that the presence of nanosilica improved the loading efficiency
as well as slowed down the release of entrapped drug.
15.4.1.2b. Chitosan: Chitosan is a mainly used biopolymer as a support because of the presence
of amino groups on its surface. Earlier chitosan microspheres were used for the immobilization
of nuclease (Shi et al. 2011). Glucose isomerase enzyme was also immobilized on chitosan beads
(Cahyaningrum et al. 2014). It was observed earlier that support-comprising chitosan in combination
with alginate showed a lesser leaching effect compared to alginate alone (Betigeri and Neau 2002).
15.4.1.2c. Cellulose and Pectin: Cellulose is an abundantly available natural polymer and it has
been extensively used for the immobilization of various enzymes such as, penicillin G acylase,
α-amylase and tyrosinase (Mislovicova et al. 2004, Namdeo and Bajpai 2009, Labus et al. 2011).
Pectin in combination with chitin and calcium alginate have improved the thermal, denaturant
resistance and biocatalytic properties of entrapped enzymes due to formation of highly stable
polyelectrolyte complexes (Gomez et al. 2006, Satar et al. 2008).
their applicability for various practical applications. Availability of silanol groups on silica’s
surface helps in attachment of biomolecules to develop biohybrids and it could be also utilized for
functionalization using chemical agents (Zucca and Sanjust 2014).
formation of biogenic silica. Silaffins are polypeptides rich in lysine and serine residues with a high
degree of post-translation modification (Kroger 2007). Sponges are the another major biosilica
producing organism and silicatein isoforms have been identified in siliceous sponge species which
play role in synthesis of biosilica (Mishra 2019). Colloidal silica nanoparticles are commercially
available and are often used for research and development purposes. One of the widely used
commercially available colloidal silica is LUDOX® colloidal silica which is aqueous dispersions of
silica nanoparticles in the nanometre size range. In India, various grades of silica nanoparticle in the
nm size range are commercially supplied by Visa chemical industries.
In 1968, a pioneering work was reported by Stober et al. (1968) for controlled synthesis of
monodispersed silica and spherical particles. This method offers advantages over acid-catalyzed
systems as using this monodispersed spherical silica particles can be prepared. However, in the
sol-gel process, production of alcohol as a by-product is one of the limitations when biological
components are going to be immobilized. Thus, there is a need to find suitable methods for
associating bio-components with silica nanoparticles to develop biohybrids and to improve the
practical applicability of associated bio-components.
Coaxial electrospinning was also applied to develop biohybrid fibres with a core-shell structure
(Letnik et al. 2015). This has a water-soluble polymer containing aqueous core which maintains the
survival and proliferation of yeast cells and an insoluble polymer shell to provide physical stability.
engineered living bacteria was associated with hydrogel-elastomer hybrid matrix and support has
helped in growth of cells. A biohybrid comprising Silica-Sphingomonas sp. cells were immobilized
and used as a biosensor for detection of methyl parathion pesticide (Mishra et al. 2017). This
biosensor could detect methyl parathion in the linear range of 0.1 – 1 ppm concentration.
As reported, Rhodobacter sphaeroides reaction centre is a robust and tractable membrane
protein which has potential for technological applications and lead to develop biohybrid devices for
various applications like biosensing (Swainsbury et al. 2014). This study evaluated the viability of
using a tiny working electrode attached to a reaction centre to create a photocurrent as the basis for
a biosensor for classes of herbicides that are widely used to control weeds in important agricultural
crops. The triazides atrazine and terbutryn suppressed photocurrent production in a concentration-
dependent manner, but nitrile or phenylurea herbicides did not. The kinetics of charge recombination
in photo-oxidized reaction centres in solution were measured, and the results revealed the same
selectivity of response. The limits of detection were calculated to be about 8 nM and 50 nM,
respectively, as indicated in the Table 15.1. Titrations of the reaction centre photocurrents produced
half maximum inhibitory doses of 2.1 mM and 208 nM for atrazine and terbutryn, respectively.
Without the need for model-dependent kinetic analysis of the signal used for detection or the use
of unreasonably complex instrumentation, photocurrent attenuation provided a direct measure of
herbicide concentration. It also offered the possibility of using protein-engineering to increase the
selectivity and sensitivity of herbicide against the Rbasphaeroides reaction centre.
In the urban and rural regions of Campo Verde and Lucas do Rio Verde Cities, Mato Grosso
State, Brazil, Table 15.1 offers information on the prevalence of the pesticides atrazine, chlorpyrifos,
a-endosulfan, β-endosulfan, flutriafol, malathion and metolachlor in water bodies (Nogueira et al.
2012). In these significant grain-producing regions, samples of the surface, rain and groundwater
were taken throughout the wet and dry seasons of 2007 and 2008. The results showed that rainwater
had a greater variety of chemicals and a higher frequency of detection than surface and groundwater
samples. Some surface and groundwater samples contained concentrations of atrazine, endosulfan
and malathion that were higher than those prescribed by Brazilian regulations, and some samples
contained a higher amount of degraded product of deisopropylatrazine and endosulfansulfate
in comparison to their parent compounds. The results highlight the fragility of the local water
infrastructure and the danger of pesticide contamination in significant headwater streams.
Law constant; eKOC: soil organic carbon sorption coefficient; ft1/2: half-live; gSw: water solubility -: metabolite; Pesticide Properties Database (PPDB). (Agriculture & Environment Research
Unit, University of Hertfordshire, PPDB: The Pesticide Properties Database, http://sitem.herts.ac.uk/aeru/footprint/en/index.htm accessed in June 2012.)
274 Bioremediation for Sustainable Environmental Cleanup
Along with sol-gel, spray drying and other advanced techniques like 3D bioprinting and
microfluidics could be applied for developing efficient biohybrids. Using these techniques,
biohybrids in various morphologies such as microparticles, sheet, fibre and scaffold could be
understood. Depending on their morphology, biohybrids could be used for various applications like
drug/gene delivery, biosensing, biomedical and environmental remediation.
Though a lot of advances have been made in the field of synthesis and application of biohybrids,
there are yet some challenges involved in synthesizing biohybrids in large quantities with accurate
precision and design, and understanding applications of biohybrids for constant monitoring and
applying them for continuous reactors. For this purpose, there is need to work in an interdisciplinary
way, where scientists from different areas like material science, biological sciences, chemical
sciences could come together. Regarding biocomponents, there is need for developing engineered
biomolecules with high biological activity, which could be associated with suitable supports. Till
today, as support organic and nanoparticle materials have been used , now there is need to develop
biogenic supports with extraordinary characteristics like intricate design of the surface structure and
chemical moieties. There is also a need to develop suitable fabricating process/techniques which
have precise control over operational parameters and in which the potential of both components
could be fully utilized.
In summary, the combination of biomolecules with supports enables biohybrid materials to act
as biomimetic material which results in a dynamic smart material. One can clearly see that advanced
characteristics (biocompatibility, flexibility, sensitivity towards analytes) associated with biohybrid
materials will enable them to achieve the desired results not only in biosensing and remediation but
also in several research areas.
References
Agriculture & Environment Research Unit, University of Hertfordshire, PPDB: The Pesticide Properties Database,
2012. http://sitem.herts.ac.uk/aeru/footprint/en/index.htm accessed in June 2012.
Ana, S. S., L. F. Santos, M. C. Mendes and J. F. Mano. 2020. Multi-layer pre-vascularized magnetic cell sheets for
bone regeneration. Biomater. 231: 119664.
Ashly, P. C., M. J. Joseph and P. V. Mohanan. 2011. Activity of diastase a-amylase immobilized on polyanilines
(PANIs). Food Chem. 127: 1808–1813.
Avnir, D., S. Braun, O. Lev and M. Ottolenghi. 1994. Enzymes and other proteins entrapped in sol-gel Materials.
Chem. Mater. 6: 1605.
Basso, A., P. Braiuca, S. Cantone, C. Ebert, P. Linda, P. Spizzo, P. Caimi, U. Hanefeld, G. Degrassi and L. Gardossi.
2007. In silico analysis of enzyme surface and glycosylation effect as a tool for efficient covalent immobilisation
of CalB and PGA on Sepabeads. Adv. Synth. Catal. 349: 877–886.
Betigeri, S. S. and S. H. Neau. 2002. Immobilization of lipase using hydrophilic polymers in the form of hydrogel
beads. Biomater. 51: 3627.
Bergna, H. E. and W. O. Roberts. 2006. Colloidal science in colloidal silica fundamentals and applications.
pp. 575–588. CRC Press: Boca Raton.
Boldridge, D. 2010. Morphological characterization of fumed silica aggregates. Aerosol. Sci. Technol. 44: 182.
Cahyaningrum, S. E., N. Herdyastusi and D. K. Maharani. 2014. Immobilization of glucose isomerase in surface-
modified chitosan gel beads. Res. J. Pharm. Biol. Chem. Sci. 5: 104.
Capulli, A. K., M. Y. Emmert, F. S. Pasqualini, D. Kehl, E. Caliskan, J. U. Lind, S. P. Sheehy, S. J. Park, S. Ahn, B.
Weber, J. A. Goss, S. P. Hoerstrup, K. K. Parker. 2017. JetValve: rapid manufacturing of biohybrid scaffolds
for biomimetic heart valve replacement. Biomaterials 133: 229–241.
Chen, L., J. Hu, Z. Fang, Y. Qi and R. Richards. 2011. Gold nanoparticles intercalated into the walls of mesoporous
silica as a versatile redox catalyst. Ind. Eng. Chem. Res. 50: 13642.
Chen, Xi, L. Mahadevan, A. Driks and O. Sahin. 2014. Bacillus spores as building blocks for stimuli-responsive
materials and nanogenerators. Nat. Nanotechnol. 9(2): 137.
Chang-Hyung., C., H. Wang, H. Lee, J. H. Kim, L. Zhang, A. Mao, D. J. Mooney and D. A. Weitz. 2016. One-step
generation of cell-laden microgels using double emulsion drops with a sacrificial ultra-thin oil shell. Lab Chip.
16(9): 1549.
Biohybrids for Environmental Remediation and Biosensing 275
Choi, C.-H., H. Wang, H. Lee, J. H. Kim, L. Zhang, A. Mao, D. J. Mooney and D. A. Weitz. 2016. One-step generation
of cell-laden microgels using double emulsion drops with a sacrificial ultra-thin oil shell. Lab Chip 16 (9):
1549–1555.
Chong, M. N., B. Jin, C. W. K. Chow and C. Saint. 2010. Recent developments in photocatalytic water treatment
technology: a review. Water Res. 44: 2997–3027. doi: 10.1016/j.watres.2010.02.039.
Compaan, A. M.., K. Song, W. Chai and Y. Huang. 2020. Cross-linkable microgel composite matrix bath for embedded
bioprinting of perfusable tissue constructs and sculpting of solid objects. ACS Appl. Mater. Interfaces. 12(7):
7855.
Coradin, T., N. Nassif and J. Livage. 2003. Silica-alginate composites for microencapsulation. Appl. Microbiol.
Biotechnol. 61: 429.
Davydov, V. Y. 2000. Adsorption on silica surfaces; E. Papirer [Ed.]. CRC Press, Taylor & Francis Group: Santa
Barbara, CA, USA, 90: 63.
Douglas, G. B., M. Lurling and B. M. Spears. 2016. Assessment of changes in potential nutrient limitation in an
impounded river after application of lanthanum-modified bentonite. Water Res. 97: 47–54. doi: 10.1016/j.
watres. 2016.02.005.
Foresti, M. L., G. Valle, R. Bonetto, M. L. Ferreira and L. E. Briand. 2010. FTIR, SEM and fractal dimension
characterization of lipase B from Candida antarctica immobilized onto titania at selected conditions. Appl.
Surf. Sci. 256: 1624–1635.
Gabriel, U., L. Charlet, C. W. Schlapfer, J. C. Vial, A. Brachmann and G. Geipel. 2001. Uranyl surface speciation on
silica particles studied by time-resolved laser-induced fluorescence spectroscopy. J. Colloid. Interf. Sci. 239:
358.
Gerber, L. C., F. M. Kbehler, R. N. Grass and J. S. Wendelin. 2012. Incorporating microorganisms into polymer layers
provides bioinspired functional living materials. Proc. Natl. Acad. Sci. U. S. A 109(1): 90.
Gomez, L., H. L. Ramırez, A. Neira-Carrillo and R. Villalonga. 2006. Polyelectrolyte complex formation mediated
immobilization of chitosan–invertase neoglycoconjugate on pectin-coated chitin. BioprocBiosyst. Eng.
28: 387.
Headen, D. M., G. Aubry, H. Lu and A. J. Garcia. 2014. Microfluidic-based generation of size-controlled,
biofunctionalized synthetic polymer microgels for cell encapsulation. Adv. Mater. 26(19): 3003.
Huang, J., S. Liu, C. Zhang, X. Wang, J. Pu, F. Ba, S. Xue, H. Ye, T. Zhao, K. Li, Y. Wang, J. Zhang, L. Wang,
C. Fan, T. K. Lu and C. Zhong. 2019. Programmable and printable Bacillus subtilis biofilms as engineered
living materials. Nat. Chem. Biol. 15(1): 34.
Hyde, E. D. E. R., A. Seyfaee, F. Neville and R. Moreno-Atanasio. 2016. Colloidal silica particle synthesis and future
industrial manufacturing pathways: a review. Ind. Eng. Chem. Res. 55: 8891.
Jamal, M., S. S. Kadam, R. Xiao, F. Jivan, T. M. Onn, R. Fernandes, T. D. Nguyen and D. H. Gracias. 2013. Bio
origami hydrogel scaffolds composed of photocrosslinked PEG bilayers. Adv. Health. Mater. 2(8): 1142.
Jimenez-Hamann, M. C. and B. A. Saville. 1996. Enhancement of tyrosinase stability by immobilization on Nylon 66.
Food Bioprod. Process Trans. Inst. Chem. Eng. C. 74: 47–52.
Jin, Y., W. Chai and Y. Huang. 2018. Fabrication of stand-alone cell-laden collagen vascular network scaffolds using
fugitive pattern-based printing-then-casting approach, ACS Appl. Mater. Interfaces. 10(34): 28361.
Kamperman, T., S. Henke, A. van den Berg, S. R. Shin, A. Tamayol, A. Khademhosseini, M. Karperien and J. Leijten.
2017. Single cell microgel based modular bioinks for uncoupled cellular micro- and macroenvironments. Adv.
Health. Mater. 6(3): 1600913.
Kamperman, T., S. Henke, W. C. Visser, M. Karperian and J. Leijten. 2017. Centering single cells in microgels via
delayed crosslinking supports long-term 3D culture by preventing cell escape. Small. 13(22): 1603711.
Katiyar, A., S. Yadav, P. G. Smirniotis and N. G. Pinto. 2006. Synthesis of ordered large pore SBA-15 spherical
particles for adsorption of biomolecules. J. Chromato. A. 1122: 1-2(13).
Kazy, S. K., S. F. D’Souza and P. Sar. 2009. Uranium and thorium sequestration by a Pseudomonas sp., Mechanism
and chemical characterization. J. Hazard. Mater. 163: 65.
Kim, S. F., R. B. John and L. W. James. 2000. Encapsulation of sulfate-reducing bacteria in a silica host. J. Mater.
Chem. 10: 1099.
Kim, J. K., J. K. Park and H. K. Kim. 2004. Synthesis and characterization of nanoporous silica support for enzyme
immobilization. Colloids Surf., A. 241: 113.
Kroger, N. 2007. Prescribing diatom morphology: toward genetic engineering of biological nanomaterials. Curr.
Opin. Chem. Biol. 11: 662.
Labus, K., A. Turek, J. Liesiene and J. Bryjak. 2011. Efficient Agaricusbisporus tyrosinase immobilization on
cellulose-based carriers. Biochem. Eng. J. 56: 232.
Lahiri, S., A. Mishra, D. Mandal, R. L. Bhardwaj and P. R. Gogate. 2021. Sonochemical recovery of uranium from
nanosilica-based sorbent and its biohybrid. Ultrasonics Sonochemistry. 76: 105667.
276 Bioremediation for Sustainable Environmental Cleanup
Leng, L., A. McAllister, B. Zhang, M. Radisic and A. Günther. 2012. Mosaic hydrogels: one-step formation of
multiscale soft materials. Adv. Mater. 24(27): 3650.
Letnik, I., R. Avrahami, J. S. Rokem, A. Greiner, E. Zussman and C. Greenblatt. 2015. Living composites of
electrospun yeast cells for bioremediation and ethanol production. Biomacromolecules. 16(10): 3322.
Li, R., L. Morrison, G. Collins, A. Li and X. Zhan. 2016. Simultaneous nitrate and phosphate removal from wastewater
lacking organic matter through microbial oxidation of pyrrhotite coupled to nitrate reduction. Water Res. 96:
32–41. doi: 10.1016/j.watres.2016.03.034.
Lienemann, P. S., T. Rossow, A. S. Mao, Q. Vallmajo-Martin, M. Ehrbar and D. J. Mooney. 2017. Single cell-laden
protease-sensitive microniches for long term culture in 3D. Lab Chip. 17(4): 727.
Liu, X., T. Tzu-Chieh, E. Tham, H. Yuk, S. Lin, T. K. Lu and X. Zhao. 2017. Stretchable living materials and devices
with hydrogel–elastomer hybrids hosting programmed cells. Proc. Natl. Acad. Sci. U. S. A. 114(9): 2200.
Liu, B. F., Y. R. Jin, Q. F. Cui, G. J. Xie, Y. N. Wu and N. Q. Ren. 2015. Photofermentation hydrogen production
by Rhodopseudomonasspnov strain A7 isolated from the sludge in a bioreactor. Int. J. Hydrogen Energy.
40: 8661–8668.
Mao, A. S., J. W. Shin, S. Utech, H. Wang, O. Uzun, W. Li, M. Cooper, Y. Hu, L. Zhang, D. A. Weitz and D. J.
Mooney. 2017. Deterministic encapsulation of single cells in thin tunable microgels for niche modelling and
therapeutic delivery. Nat. Mater. 16(2): 236.
Melo, J. S., A. Tripathi, J. Kumar, A. Mishra, S. Bhanu and K. Bhainsa. 2020. Immobilization: then and now.
pp. 1–84. In: A. Tripathi and J. S. Melo [Eds.]. I Immobilization Strategies. Gels Horizons: From Science to
Smart Materials. Singapore: Springer.
Metilda, J. P., M. Gladis and T. P. Rao. 2005. Catechol functionalized aminopropyl silica gel: synthesis, characterization
and preconcentrative separation of uranium (VI) from thorium (IV), Radiochim. Acta. 93: 219.
Mishra, A., J. S. Melo, D. Sen and S. F. D’Souza. 2014. Evaporation induced self assembled microstructures of silica
nanoparticles and Streptococcus lactis cells as sorbent for uranium (VI). J. Colloid Interface Sci. 414: 33–40.
Mishra, A., J. Kumar and J. S. Melo. 2017. An optical microplate biosensor for the detection of methyl parathion
pesticide using a bio-hybrid of Sphingomonas sp. cells-silica nanoparticles. Biosensor and Bioelectronics. 87:
332–338.
Mishra, A. 2019. Synthesis, characterization and applications of silica based biohybrid materials, Ph.D. Thesis, Homi
Bhabha National Institute, Mumbai, India.
Mishra, A., J. S. Melo, A. Agrawal, Y. Kashyap and D. Sen. 2020a. Preparation and application of silica nanoparticles-
Ocimumbasilicum seeds bio-hybrid for the efficient immobilization of invertase enzyme. Colloids Surface B.
Biointerfaces. 188: 110796.
Mishra, A., J. Kumar and J. S. Melo. 2020b. Silica based bio-hybrid materials and their relevance to bionanotechnology.
Austin J. Plant Biol. 6(1): 1024–1028.
Mishra, A., V. K. Pandey, B. S. Shankar and J. S. Melo. 2021a. Spray drying as an efficient route for synthesis of
silica nanoparticles-sodium alginate biohybrid drug carrier of doxorubicin. Colloids Surface B: Biointerfaces.
197: 111445.
Mishra, A., J. Kumar, J. S. Melo and B. P. Sandaka. 2021b. Progressive development in biosensors for detection of
dichlorvos pesticide: a review. J. Environ. Chem. Eng. 9(2): 105067.
Mishra, A., S. Mukundan and J. Kumar. 2022. An overview of metal-organic frameworks for detection of pesticides. In:
Ram K. Gupta, Tahir Rasheed, Tuan Anh Nguyen and Muhammad Bilal [Eds.]. I Metal-Organic Frameworks-
Based Hybrid Materials for Environmental Sensing and Monitoring. CRC Press Taylor & Francis, USA, P. 7,
eBook ISBN 9781003188148. doi.10.1201/9781003188148-21.
Mislovicova, D., J. Masarova, A. Vikartovska, P. Germeiner and E. Michalkova. 2004. Biospecific immobilization
of mannan–penicillin G acylase neoglycoenzyme on Concanavalin A-bead cellulose. J. Biotechnol. 110: 11.
Mukundan, S., J. S. Melo, D. Sen and J. Bahadur. 2020. Enhancement in β-galactosidase activity of Streptococcus
lactis cells by entrapping in microcapsules comprising of correlated silica nanoparticles. Colloids Surface B:
Biointerfaces. 195: 111245.
Namdeo, M. and S. K. Bajpai. 2009. Immobilization of a-amylase onto cellulose-coated magnetite (CCM)
nanoparticles and preliminary starch degradation study. J. Mol. Catal. B-Enzym. 59: 134.
Nandiyanto, A. B. D., S.-G. Kim, F. Iskandar and K. Okuyama. 2009. Synthesis of silica nanoparticles with nanometer-
size controllable mesopores and outer diameters. Microporous and Mesoporous Mater. 20(3): 447.
Neal, D., M. S. Sakar, L.-L. S. Ong and H. H. Asada. 2014. Formation of elongated fascicle-inspired 3D tissues
consisting of high-density, aligned cells using sacrificial outer molding. Lab Chip. 14(11): 1907.
Nguyen, P. Q., N.-M. D. Courchesne, A. Duraj-Thatte, P. Praveschotinunt and N. S. Joshi. 2018. Engineered living
materials: prospects and challenges for using biological systems to direct the assembly of smart materials. Adv.
Mater. 30(19): e1704847.
Biohybrids for Environmental Remediation and Biosensing 277
Nogueira, E. N., E. F. G. C. Dores, A. A. Pinto, R. S. Amorim, M. L. Ribeiro and C. Lourencetti. 2012. Currently used
pesticides in water matrices in Central-Western Brazil. J. Braz. Chem. Soc. 23(8): 1476–1487.
Oh, S. Y., Y. D. Seo, B. Kim, I. Y. Kim and D. K. Cha. 2016. Microbial reduction of nitrate in the presence of zero
valent iron and biochar. Bioresour. Technol. 200: 891–896. doi: 10.1016/j.biortech.2015.11.021.
Onoe, H., T. Okitsu, A. Itou, M. Kato-Negishi, R. Gojo, D. Kiriya, K. Sato, S. Miura, S. Iwanaga, K. Kuribayashi-
Shigetomi, Y. T. Matsunaga, Y. Shimoyama and S. Takeuchi. 2013. Metre-long cell-laden microfibres exhibit
tissue morphologies and Functions. Nat. Mater. 12(6): 584.
Ouyang, L. H., B. Christopher, W. Sun and J. A. Burdick. 2017. A generalizable strategy for the 3D bioprinting of
hydrogels from nonviscous photo-crosslinkable inks. Adv. Mater. 29(8): 1604983.
Ouyang, L., J. P. K. Armstrong, M. Salmeron-Sanchez and M. M. Stevens. 2020. Assembling living building blocks
to engineer complex tissues. Adv. Funct. Mater. 30(26): 1909009.
Pattanawanidchai, S., S. Loykulnant, P. Sae-oui, N. Maneevas and C. Sirisinha. 2014. Development of eco-friendly
coupling agent for precipitated silica filled natural rubber compounds. Polym. Test. 34: 58.
Prasertsri, S. and N. Rattanasom. 2012. Fumed and precipitated silica reinforced natural rubber composites prepared
from latex system: mechanical and dynamic properties. Polym. Test. 31: 593.
Premkumar, J. R., O. Lev, R. Rosen and S. Belkin. 2001. Encapsulation of luminous recombinant E. coli in sol-gel
silicate films. Adv. Mater. 13(23): 1773.
Pu, J., Y. Liu, J. Zhang, B. An, Y. Li, X. Wang, K. Din, C. Qin, K. Li, M. Cui, S. Liu, Y. Huang, Y. Wang, Y. Lv,
J. Huang, Z. Cui, Z. Suwen and C. Zhong. 2020. Virus disinfection from environmental water sources using
living engineered biofilm materials. Adv. Sci. 7(14): 1903558.
Puig, M., L. Cabedo, J. J. Gracenea, A. Jiménez-Morales, J. Gámez-Pérez and J. J. Suay. 2014. Adhesion enhancement
of powder coatings on galvanised steel by addition of organo-modified silica particles. Prog. Org. Coat.
77: 1309.
Rahman, I. A. and V. Padavettan. 2012. Synthesis of silica nanoparticles by sol-gel: size-dependent properties, surface
modifications and applications in silica-polymer nanocomposites—a review. J. Nanomater. 1.
Rajanna, S. K., D. Kumar, M. Vinjamur and M. Mukhopadhyay. 2015. Silica aerogel microparticles from rice husk
ash for drug delivery. Ind. Eng. Chem. Res. 54: 949.
Reshmi, R., G. Sanjay and S. Sugunan. 2007. Immobilization of a-amylase on zirconia: a heterogeneous biocatalyst
for starch hydrolysis. Catal. Commun. 8: 393.
Rimola, A., D. Costa, M. Sodupe, J.-F. Lambert and P. Ugliengo. 2013. Silica surface features and their role in the
adsorption of biomolecules: computational modelling and experiments. Chem. Rev. 113: 4216.
Rivera-Tarazona, L. K., Z. T. Campbell and T. H. Ware. 2021. Stimuli-responsive engineered living materials. Soft
Matter. 17(4): 785.
Roxby, D. N., H. Rivy, C. Gong, X. Gong, Z. Yuan, G. E. Chang and Y. C. Chen. 2020. Microalgae living sensor
for metal ion detection with nanocavity enhanced photoelectrochemistry. Biosens. Bioelectron. 165: 112420.
Sankaran, S., J. Becker, C. Wittman and A. dCampo. 2019. Optoregulated drug release from an engineered living
material: self-replenishing drug depots for long-term, light-regulated delivery. Small. 15(5): e1804717.
Sar, P., S. K. Kazy and S. F. D’Souza. 2004. Radionuclide remediation using a bacterial biosorbent. Int. Biodeter.
Biodegr. 54: 193.
Satar, R., M. Matto and Q. Husain. 2008. Studies on calcium alginate—pectin gel entrapped concanavalin A–bitter
gourd (Momordica charantia) peroxidase complex. J. Sci. Ind. Res. India. 67: 609.
Shi, L. E., Z. X. Tang, Y. Yi, J. S. Chen, W. Y. Xiong and G. Q. Ying. 2011. Immobilization of nuclease p1 on chitosan
micro-spheres. Chem. Biochem. Eng. Q. 25: 83.
Shukla, P., A. Mishra, S. Manivannan, J. S. Melo and D. Mandal. 2020. Parametric optimization for adsorption of
mercury (II) using self-assembled bio-hybrid. J. Environ. Chem. Eng. 8(3): 103725.
Shukla, P., A. Mishra, S. Manivanna and D. Mandal. 2022. Metal-Organic-Frames (MOFs) based electrochemical
sensors for sensing heavy metal contaminated liquid effluents: a review. Nanoarchitectonics. 3(2): 46–60.
Shukla, V. Y., D. R. Tipre and S. R. Dave. 2014. Optimization of chromium (VI) detoxification by Pseudomonas
aeruginosa and its application for treatment of industrial waste and contaminated soil. Bioremediat. J. 18:
128–135. doi: 10.1080/10889868.2013.834872.
Son, J., C. Y. Bae and J.-K. Park. 2016. Freestanding stacked mesh-like hydrogel sheets enable the creation of complex
macroscale cellular scaffolds. Biotechnol. J. 11(4): 585.
Sun, Z., C. Bai, S. Zheng, X. Yang and R. L. Frost. 2013. A comparative study of different porous silica minerals
supported TiO2 catalysts. Appl. Catal. A. 458: 103.
Stöber, W., A. Fink and E. Bohn. 1968. Controlled growth of monodisperse silica spheres in the micron size range. J.
Colloid. Inter. Sci. 26(1): 62.
278 Bioremediation for Sustainable Environmental Cleanup
Swainsbury, D. J. K., V. M. Friebe, R. N. Frese and M. R. Jones. 2014. Evaluation of a biohybrid photoelectrochemical
cell employing the purple bacterial reaction centre as a biosensor for herbicides. Biosensors and Bioelectronics.
58: 172–178.
Taikum, O., R. Friehmelt and M. Scholz. 2010. The last 100 years of fumed silica in rubber reinforcement. Rubber
World. 242: 35.
Tang, F., L. Li and D. Chen. 2012. Mesoporous silica nanoparticles: synthesis, biocompatibility and drug delivery.
Adv. Mater. 24: 1504.
Tian, D., P. Dubois and R. Jérôme. 1997. Biodegradable and biocompatible inorganic–organic hybrid materials. I.
Synthesis and characterization. J. Poly. Sci. Part A: Poly. Chem. 35(11): 2295.
Valenti, G., R. Rampazzo, S. Bonacchi, L. Petrizza, M. Marcaccio, M. Montalti, L. Prodi and F. Paolucci. 2016.
Variable doping induces mechanism swapping in electrogenerated chemiluminescence of Ru(bpy)32+ core−
shell silica nanoparticles. J. Am. Chem. Soc. 138(49): 15935.
Vassilev, N., M. Fenice, F. Federici and R. Azcon. 1997. Olive mill waste water treatment by immobilized cells of
Aspergillus niger and its enrichment with soluble phosphate. Process. Biochem. 32: 617–620. doi: 10.1016/
S0032-9592(97) 00024-1.
Vartiainen, J., M. Rättö and S. Paulussen. 2005. Antimicrobial activity of glucose oxidase-immobilized plasma-
activated polypropylene films. Packag. Technol. Sci. 18: 243–251.
Wang, J.-S., X.-J. Hu, Y.-G. Liu, S.-B. Xie and Z.-L. Bao. 2010. Biosorption of uranium (VI) by immobilized
Aspergillus fumigatus beads. J. Environ. Radioactiv. 101: 504.
Wang, K., P. Liu, Y. Ye, J. Li, W. Zhao and X. Huang. 2014. Fabrication of a novel laccase biosensor based on silica
nanoparticles modified with phytic acid for sensitive detection of dopamine. Sens. Actuators B. 197: 292.
Wang, W., W. Zhou, J. Li, Hao, D. Z. Su and G. Ma. 2015. Comparison of covalent and physical immobilization of
lipase in gigaporous polymeric microspheres. Bioprocess Biosys. Eng. 38: 2107.
Wang, X., Y. Yu, C. Yang, C. Shao, K. L. Shi, Shang, F. Ye and Y. Zhao. 2021. Microfluidic 3D printing responsive
scaffolds with biomimetic enrichment channels for bone regeneration. Adv. Funct. Mater. 31(40): 2105190.
Wang, C., Z. Zhang, J. Wang, Q. Wang and L. Shang. 2022. Biohybrid materials: structure design and biomedical
applications. Materials Today Bio. 16: 100352.
Xie W., L. Hu and X. Yang. 2015. Basic ionic liquid supported on mesoporous SBA-15 silica as an efficient
heterogeneous catalyst for biodiesel production. Ind. Eng. Chem. Res. 54: 1505.
Xu, H., M. Medina-Sánchez, V. Magdanz, L. Schwarz, F. Hebenstreit and O. G. Schmidt. 2018. Sperm-hybrid
micromotor for targeted drug delivery. ACS Nano. 12(1): 327.
Yang, S. H., J. Choi, L. Palanikumar, E. S. Choi, J. Lee, J. Kim, I. S. Choi and J.-H. Ryu. 2015. Cytocompatible in situ
cross-linking of degradable LbL films based on thiol–exchange reaction. Chem. Sci. 6(8): 4698.
Yang, Y., H. Yang, M. Yang, Y. Liu, G. Shen and R. Yu. 2004. Amperometric glucose biosensor based on a surface
treated nanoporous ZrO2/Chitosan composite film as immobilization matrix. Analytica. Chimica. Acta.
525(2): 213.
Yang, Z., S. Si and C. Zhang. 2008. Study on the activity and stability of urease immobilized ontonanoporous alumina
membranes. Microporous Mesoporous Mater. 111: 359.
Zdarta, J., A. S. Meyer, T. Jesionowski and M. Pinelo. 2018. A general overview of support materials for enzyme
immobilization: characteristics, properties, practical utility. Catalysts. 8: 92.
Zhang, Y., K. Yan, F. Ji and L. Zhang. 2018. Enhanced removal of toxic heavy metals using swarming biohybrid
adsorbents. Adv. Funct. Mater. 1806340.
Zhao, W., Y. Fang, Q. Zhu, K. Wang, M. Liu, X. Huang and J. Shen. 2013. A novel glucose biosensor based on
phosphonic acid-functionalized silica nanoparticles for sensitive detection of glucose in real samples.
Electrochim. Acta. 89: 278.
Zhu, L. J., L. P. Zhu, J. H. Jiang, Z. Yi, Y. F. Zhao, B. K. Zhu and Y. Y. Xu. 2014. Hydrophilic and anti-fouling
polyethersulfone ultrafiltration membranes with poly(2-hydroxyethyl methacrylate) grafted silica nanoparticles
as additive. J. Membr. Sci. 451: 157.
Zhu G., Q. Wang, J. Yin, Z. Li, P. Zhang, B. Ren, G. Fan and P. Wan. 2016. Toward a better understanding of
coagulation for dissolved organic nitrogen using polymeric zinc-iron-phosphate coagulant. Water Res. 100:
201–210. doi: 10.1016/j.watres. 2016.05.035.
Zhu, W., J. Guo, S. Amini, Y. Ju, J. O. Agola, A. Zimpel, J. Shang, A. Noureddine, F. Caruso, S. Wuttke,
J. G. Croissant and C. J. Brinker. 2019. SupraCells: living mammalian cells protected within functional
modular nanoparticle-based exoskeletons. Adv. Mater. 31(25): e1900545.
Zucca, P. and E. Sanjust. 2014. Inorganic materials as supports for covalent enzyme immobilization: methods and
mechanisms. Mol. 19: 14139.
Chapter 16
Remediation for Heavy Metal
Contamination
A Nanotechnological Approach
Rubina Khanam, Amaresh Kumar Nayak* and
Dibyendu Chatterjee
16.1 Introduction
Pollution is defined as the presence of undesirable chemical objects that obstruct natural processes
or have negative consequences for living beings and the environment. Pollution is increasing
at an alarming rate as a result of industrialization and the massive rise in population that leads
to rising urbanization. Environmental pollution identification, treatment and prevention is a
critical step toward long-term environmental sustainability. Environment sustainability refers to
the responsible and justifiable relationship between humans and the environment, as well as the
intelligent use of resources, ensuring environmental safety for current and future generations.
Economic and environmental sustainability are inextricably linked (Fajardo et al. 2020). A lot of
work is being done right now to discover and create persuasive and dependable ways for degrading
or transforming environmental contaminants of concern. Nanoremediation is a groundbreaking
remediation technique that employs nanomaterials having high surface: volume ratio, low reduction
potential and quantum confinement making them efficient for the detoxification and alteration of
hazardous recalcitrant pollutants in the system (Fajardo et al. 2020). In particular, when compared
to standard remediation procedures (viz., chemical oxidation, thermal decomposition and solvent
co-flushing), the use of nanomaterials in environmental remediation has gained a lot of attention.
Nanoremediation methods have drawn a lot of interest because of their unique qualities, such as cost-
effectiveness, sensitivity, superior electrical properties, high surface area and improved catalytic
properties (Shafi et al. 2021). These techniques have the potential to give long-term solutions
to environmental pollution issues, while also reducing the cost of cleaning (Shafi et al. 2021).
Nanostructure-based technologies have the potential to reduce not only the overall costs of cleaning
up large-scale contaminated areas but also to reduce clean-up time, minimize the need for polluted
material treatment and disposal, and decrease pollutant concentrations to near zero-all in-situ
(Corsi et al. 2018). Nanoremediation involves the applications of nano-sized metal and bimetallic
particles, nanodots, carbon nanotubes and nanocomposites, etc., for breaking down contaminants
(Shafi et al. 2021). These nanomaterials in the form of sensors, catalysts and adsorbents ensure
rapid detection and immediate detoxification of pollutants such as heavy metals and metalloids from
contaminated land sites (Figure 16.1). This chapter summarizes the application of nanotechnology
(1) to detect and quantify trace pollutants in the environment. (2) to decontaminate the environment
using nano adsorbents, nanocatalysts, nano clay composites, etc.
Nanoseners
Figure 16.1. Application nanoremediation techniques for reducing heavy metal contamination in the environment.
Figure. 16.1 Application Nnanoremediation techniques for reducing heavy metal
detection that can selectively bind to As3+ ions via the As-O link (Han et al. 2010). SERS provides
information on the existence of a chemical element as well as its chemical composition, which is
significant in metal ion toxicological studies since different heavy metals complexes have different
degrees of toxicity in humans and animals.
Table 16.4. Chemically synthesized nanomaterials used for detecting heavy metals.
16.4.1 Nano-photocatalyst
Photocatalysis is a photochemical reaction based on electron/hole pair redox interactions when
exposed to light (Danish et al. 2021). Photocatalysis has the potential to degrade contaminants
that are difficult to decompose. The photocatalytic process is regarded as an effective remediation
method for converting harmful chemicals into environmentally friendly products. In the presence
of abundant solar radiation, photocatalysts accelerate chemical reactions. In general, photocatalysis
is a redox process in which holes are created in the Valence Band (VB) of a photocatalyst and
electrons are generated in the Conduction Band (CB), resulting in the emergence of highly
energetic and reactive species such as hydroxyl (OH–) and superoxide radicals (O2–). Due to the
production of extremely energetic radicals that function as potent oxidizing agents, photocatalysis
serves as a fruitful technique for the removal of hazardous chemical compounds (Ajmal et al.
2014). Nanomaterial photocatalysis for environmental cleanup has been the subject of significant
research for the past decade. To cleanse soil, purify the air and detoxify wastewater, a wide
range of nanostructured photocatalysts have been synthesized (viz., oxides and sulfide of metals,
composite oxides, carbon derivatives, graphene-based photocatalysts, dendrimers and polymeric
nanocomposites (Danish et al. 2021). In comparison to conventional photocatalysts, graphene-based
photocatalysts have been found to enhance the activity due to their large surface area, nanosize
and greater electronic motions. Oxide-based nanomaterials, such as Fe3O4, TiO2, ZnO and their
composites, are also excellent catalysts for the removal of heavy metals. Several researches
have reported the use of these oxides in environmental remediation, notably in photocatalytic
degradation of organic contaminants (Czech et al. 2020, Sabzehei et al. 2020, Masudy-Panah
et al. 2019). By reducing the distance between photon absorption sites and limiting electron-hole
(e– -h+) recombination, these nanoparticles with a large surface area and porosity have greater
photocatalytic activity. Transition metal oxides and their composites have strong photocatalytic
activity for organic pollutant photodegradation. TiO2 has demonstrated outstanding photocatalytic
destruction of contaminants due to its high resistance against photochemical, exceptional surface
qualities, microstructural features and large surface area; nanosized metal oxides are employed
in adsorption processes (Hitam et al. 2018, Rahman et al. 2011). The surface energy of metal
absorbents is improved by decreasing their size and generating more active sites on their surface for
the adsorption of pollutant molecules (Gusain et al. 2019). Immobilizing ZnO NPs on polymer
substrates is another agreeable technique in the production of modified-ZnO photocatalysts.
According to research, ZnO/polymer nanocomposite achieves the requisite photocatalytic activity
(Shirdel and Behnajady 2020). Hydrothermal synthesis and homogeneous precipitation are the
most appropriate procedures to create ZnO photocatalytic materials. A large number of studies have
demonstrated CuO’s remarkable performance in the photocatalytic breakdown of contaminants.
Chen et al. (2020) reported a 97% photodegradation of crystal violet dye utilizing monoclinic
crystalline CuO nanoparticles when exposed to visible light. Carbon nanotubes (CNTs) are also
used as innovative materials for photocatalyst due to their higher quantum efficiency and excellent
chemical stability. Gupta (2017) successfully developed ultrathin photocatalyst (SWCNTs-TiO2) for
wastewater treatment.
16.4.2 Nanoadsorbents
Heavy metal discharge from industrial, municipal, agricultural and household wastewater has
become a significant ecological concern. Over the last few decades, a new class of nano-adsorbents
has been developed to combat this rising menace. Adsorption is commonly used for heavy metal
removal due to its low cost, efficiency and simplicity. They have acquired recognition as a result
of their particular characteristics, and have demonstrated outstanding promise in the treatment
of wastewater and industrial effluents for reuse in a wide range of applications for the long-term
sustainability of the environment. Adsorption has long been known as a phenomenon in water
treatment. It is a common experience in the gaseous phase, but is used effectively in the treatment
Remediation for Heavy Metal Contamination 285
of pollutants from soil and water. The specificity for a particular hazardous chemical, the adsorption
efficiency and the benefit-cost ratio are important factors in selecting adsorbents. In the recent past,
different nanomaterials viz. carbon analogues, carbonaceous nanostructured composites, nano
magnetic materials, microporous glasses, adsorbent sieves ceramics, clay polymers, etc. have been
synthesized (Kumar and Guleria 2020). In comparison to the materials utilized historically and
commercially as adsorbents, they have been found to have a high adsorption capacity for heavy
metals from wastewater. Among nanoabsorbents, carbon nanomaterials are most extensively used
for the removal of heavy metals. The carbon nanomaterials can be classified as (i) Zero-dimensional
(having all the three dimensions less than 100 nm); for example, are fullerene and quantum dots.
(ii) One-dimensional (1 D) (having only one dimension larger than 100 nm and two dimensions
smaller than 100 nm), e.g., nanotubes of carbon and titanium. (iii) Two-dimensional nanomaterials
(with two dimensions greater than 100 nm), e.g., graphene. (iv) Three dimensional (3-D) (all
dimensions are greater than 100 nm), e.g., graphite and some nono-composites. Highly efficient
nano absorbents are described next.
Table 16.5. Removal efficiency of different metal based nano absorbents and carbon nanotube-based adsorbents.
10 nm and coated with ascorbic acid were highly efficient in removing As. These nanoadsorbentions
had the maximum adsorption capacity of 16.56 mg g–1 and 46.06 mg g–1 for As(III) and As(V),
respectively (Feng et al. 2012). Further, Fe2O4 MNPs coated with oxide shells of Mn-Co is an
excellent adsorber of Pb (II) (481.2 mg g–1), Cu (II) (386.2 mg g–1) and Cd (II) (345.5 mg g–1)
(Ma et al. 2013). Silica nanoparticles are also considered to be highly efficient in removing heavy
metals. Silica nanoparticles along with graphite oxide were used to remove Zn, Ni, Cr, Pb and Cd
by Sheet et al. (2014). When metal oxides are used as adsorbents without any supporting elements
(viz., clays or zeolites), they may have limited adsorption capacity, smaller surface area and release
hazardous metals into the environment.
(Sakulthaew et al. 2017). Recently, chitosan-based clay composite showed high adsorption of
Cu (II) in the tune of 96.0 and 99.5%, respectively, at a pH of 6.5 (Sakulthaew et al. 2017). In
addition, intercalation polymerization generated organo-bentonite and polyacrylonitrile composite
showed the removal of Cd, Zn and Cu at the tune of 52.6, 65.4 and 77.4 mg g–1, respectively
(Mukhopadhyay et al. 2020).
References
Ajmal, A., I. Majeed, R. N. Malik, H. Idriss and M. A. Nadeem. 2014. Principles and mechanisms of photocatalytic
dye degradation on TiO2 based photocatalysts: a comparative overview. Rsc. Adv. 4(70): 37003–37026.
Alijani, H. and Z. Shariatinia. 2018. Synthesis of high growth rate SWCNTs and their magnetite cobalt sulfide
nanohybrid as super-adsorbent for mercury removal. Chem. Eng. Res. Des. 129: 132–149.
Anitha, K., S. Namsani and J. K. Singh. 2015. Removal of heavy metal ions using a functionalized single-walled
carbon nanotube: a molecular dynamics study. J. Phys. Chem. A. 119(30): 8349–8358.
Bhargavi, R. J., U. Maheshwari and S. Gupta. 2015. Synthesis and use of alumina nanoparticles as an adsorbent for
the removal of Zn(II) and CBG dye from wastewater. Int. J. Ind. Chem. 6: 31–41.
Bleiman, N. and Y. G. Mishael. 2010. Selenium removal from drinking water by adsorption to chitosan–clay
composites and oxides: batch and columns tests. J. Hazard. Mater. 183(1-3): 590–595.
Boken, J. and D. Kumar. 2014. Detection of toxic metal ions in water using SiO 2@ Ag Core-Shell nanoparticles. Int.
J. Environ. Res. Dev. 4: 303–308.
Bulut, Y. and Z. Tez. 2009. Adsorption of heavy metal ions from aqueous solutions by bentonite. J. Colloid Interface
Sci. 332(1): 46–53.
Chai, F., C. Wang, T. Wang, L. Li and Z. Su. 2010. Colorimetric detection of Pb2+ using glutathione functionalized
gold nanoparticles. ACS Appl. Mater. Interface. 2(5): 1466–1470.
Chalasani, R. and S. Vasudevan. 2012. Cyclodextrin functionalized magnetic iron oxide nanocrystal: a host-carrier
for magnetic separation of non-polar molecules and arsenic from aqueous media. J. Mater. Chem. 22:
14925–14931.
Chen, K. I., B. R. Li and Y. T. Chen. 2011. Silicon nanowire field-effect transistor-based biosensors for biomedical
diagnosis and cellular recording investigation. Nano Today. 6(2): 131–154.
Remediation for Heavy Metal Contamination 289
Chen, K., X. Wang, P. Xia and J. Xie. 2020. Efficient removal of tetrabromodiphenyl ether with a Z-scheme Cu2O
(rGOTiO2) photocatalyst under sunlight irradiation. Chemosphere. 254: 126806.
Corsi, I., M. Winther-Nielsen, R. Sethi, C. Punta, C. Della Torre, G. Libralato, G. Lofrano, L. Sabatini, M. Aiello,
L. Fiordi and F. Cinuzzi. 2018. Ecofriendly nanotechnologies and nanomaterials for environmental applications:
key issue and consensus recommendations for sustainable and ecosafenanoremediation. Ecotoxicol. Environ.
Saf. 154: 237–244.
Czech, B, P. Zygmunt, Z. C. Kadirova, K. Yubuta and M. Hojamberdiev. 2020. Effective photocatalytic removal of
selected pharmaceuticals and personal care products by Elsmoreite/Tungsten Oxide@Zns Photocatalyst. J.
Environ. Manag. 270: 110870.
Danish, M. S. S., L. L. Estrella, I. M. A. Alemaida, A. Lisin, N. Moiseev, M. Ahmadi, M. Nazari, M. Wali,
H. Zaheb and T. Senjyu. 2021. Photocatalytic applications of metal oxides for sustainable environmental
remediation. Metals. 11(1): 80.
Darbha, G. K., A. Ray and P. C. Ray. 2007. Gold nanoparticle-based miniaturized nanomaterial surface energy transfer
probe for rapid and ultrasensitive detection of mercury in soil, water, and fish. Acs Nano. 31(3): 208–14.
Dehghani, M. H., M. M. Taher, A. K. Bajpai, B. Heibati, I. Tyagi, M. Asif, S. Agarwal and V. K. Gupta. 2015. Removal
of noxious Cr (VI) ions using single-walled carbon nanotubes and multi-walled carbon nanotubes. Chem. Eng.
J. 279: 344–352.
Fajardo, C., S. Sánchez-Fortún, G. Costa, M. Nande, P. Botías, J. García-Cantalejo, G. Mengs and M. Martín. 2020.
Evaluation of nanoremediation strategy in a Pb, Zn and Cd contaminated soil. Sci. Total Environ. 706: 136041.
Feng, L., M. Cao, X. Ma, Y. Zhu and C. Hu. 2012. Superparamagnetic high-surface-area Fe3O4 nanoparticles as
adsorbents for arsenic removal. J. Hazard. Mater. 217: 439–446.
Ge, Y., Z. Li , D. Xiao, P. Xiong and N. Ye. 2014. Sulfonated multi-walled carbon nanotubes for the removal of copper
(II) from aqueous solutions. J. Ind. Eng. Chem. 20(4): 1765–71.
Gich, M., C. Fernández-Sánchez, L. C. Cotet, P. Niu and A. Roig. 2013. Facile synthesis of porous bismuth–carbon
nanocomposites for the sensitive detection of heavy metals. J. Mater. Chem. A 1(37): 11410–8.
Gupta, R. K. 2017. Oil/water separation techniques: a review of recent progresses and future directions. J. Mater.
Chem. A. 5(31): 16025–16058.
Gusain, R., K. Gupta, P. Joshi and O. P. Khatri. 2019. Adsorptive removal and photocatalytic degradation of organic
pollutants using metal oxides and their composites: a comprehensive review. Adv. Colloid Interface Sci. 272:
102009.
Han, D., S. Y. Lim, B. J. Kim, L. Piao and T. D. Chung. 2010. Mercury (II) detection by SERS based on a single gold
microshell. Chem. Commun. 46(30): 5587–5589.
Hitam, C. N. C., A. A. Jalil, S. Triwahyono, A. F. A. Rahman, N. S. Hassan, N. F. Khusnun, S. F. Jamian, C. R.
Mamat, W. Nabgan and A. Ahmad. 2018. Effect of carbon-interaction on structure-photoactivity of Cu doped
amorphous TiO2 catalysts for visible-light- oriented oxidative desulphurization of dibenzothiophene. Fuel.
216: 407–417.
Kaushal, A. and S. K. Singh. 2017. Removal of heavy metals by nanoadsorbents: a review. J. Environ. Biotechnol.
Res. 6(1): 96–104.
Khanam, R., A. Kumar, A. K. Nayak, M. Shahid, R. Tripathi, S. Vijayakumar, D. Bhaduri, U. Kumar, S. Mohanty,
P. Panneerselvam and D. Chatterjee. 2020. Metal(loid)s (As, Hg, Se, Pb and Cd) in paddy soil: bioavailability
and potential risk to human health. Sci. Total Environ. 699: 134330.
Kim, T. H., J. Lee and S. Hong. 2009. Highly selective environmental nanosensors based on anomalous response of
carbon nanotube conductance to mercury ions. J. Phys. Chem. C. 113(45): 19393–19396.
Kimmel, D. W., G. LeBlanc, M. E. Meschievitz and D. E. Cliffel. 2012. Electrochemical sensors and biosensors.
Anal. Chem. 84(2): 685–707.
Kocabas-Atakli, Z. Ö. and Y. Yürüm. 2013. Synthesis and characterization of anatase nanoadsorbent and application
in removal of lead, copper and arsenic from water. Chem. Eng. J. 225: 625–635. http://dx.doi.org/10.1016/j.
cej.2013.03.106.
Konate, A., X. He, Z. Zhang, Y. Ma, P. Zhang, G. M. Alugongo and Y. Rui. 2017. Magnetic (Fe3O4) nanoparticles
reduce heavy metals uptake and mitigate their toxicity in wheat seedling. Sustain. 9: 790.
Kumar, V. and P. Guleria. 2020. Application of DNA-Nanosensor for environmental monitoring: recent advances and
perspectives. Curr. Pollut. Rep. 12: 1–21.
Li, F., J. Wang, Y. Lai, C. Wu, S. Sun, Y. He and H. Ma. 2013a. Ultrasensitive and selective detection of copper (II) and
mercury (II) ions by dye-coded silver nanoparticle-based SERS probes. Biosens. Bioelectron. 39(1): 82–87.
Li, M., S. K. Cushing, J. Zhang, J. Lankford, Z. P. Aguilar, D. Ma and N. Wu. 2012. Shape-dependent surface-
enhanced Raman scattering in gold–Raman-probe–silica sandwiched nanoparticles for biocompatible
applications. Nanotechnol. 23(11): 115501.
290 Bioremediation for Sustainable Environmental Cleanup
Li, M., S. K. Cushing, H. Liang, S. Suri, D. Ma and N. Wu. 2013b. Plasmonic nanorice antenna on triangle nanoarray
for surface-enhanced Raman scattering detection of hepatitis B virus DNA. Anal. Chem. 85(4): 2072–2078.
Li, S., Y. Li, J. Cao, J. Zhu, L. Fan and X. Li. 2014. Sulfur-doped graphene quantum dots as a novel fluorescent probe
for highly selective and sensitive detection of Fe3+. Anal. Chem. 86(20): 10201–7.
Li, J., X. Xing, J. Li, M. Shi, A. Lin, C. Xu and R. Li. 2018. Preparation of thiol-functionalized activated carbon from
sewage sludge with coal blending for heavy metal removal from contaminated water. Environl. pollut. 234:
677–683.
Liu, J., G. Lv, W. Gu, Z. Li, A. Tang and L. Mei. 2017. A novel luminescence probe based on layered double
hydroxides loaded with quantum dots for simultaneous detection of heavy metal ions in water. J. Mater. Chem.
5(20): 5024–5030.
Ma, H., H. Wang and C. Na. 2015. Microwave-assisted optimization of platinum-nickel nanoalloys for catalytic water
treatment Appl. Catal. B Environ. 163: 198–204.
Ma, Z., D. Zhao, Y. Chang, S. Xing, Y. Wu and Y. Gao. 2013. Synthesis of MnFe2 O4 and Mn–Co oxide core–shell
nanoparticles and their excellent performance for heavy metal removal. Dalt. Trans. 42(39): 14261–14267.
Maghsoudi, A. S., S. Hassani, K. Mirnia and M. Abdollahi. 2021. Recent advances in nanotechnology-based
biosensors development for detection of arsenic, lead, mercury, and cadmium. Int. J. Nanomed. 16: 803.
Masudy-Panah, S., R. Katal, N. D. Khiavi, E. Shekarian, J. Hu and X. Gong. 2019. A high-performance cupric
oxide photocatalyst with palladium light trapping nanostructures and a hole transporting layer for
photoelectrochemical hydrogen evolution. J. Mater. Chem. A 7: 22332–22345.
Mawia, A. M., S. Hui, L. Zhou, H. Li, J. Tabassum, C. Lai, J. Wang, G. Shao, X. Wei, S. Tang, J. Luo, S. Hu and
P. Hu. 2021. Inorganic arsenic toxicity and alleviation strategies in rice. J. Hazard. Mater. 408: 124751. https://
doi.org/10.1016/j.jhazmat.2020.124751.
Mukhopadhyay, R., D. Bhaduri, B. Sarkar, R. Rusmin, D. Hou, R. Khanam et al. 2020. Clay–polymer nanocomposites:
progress and challenges for use in sustainable water treatment. J. Hazard. Mater. 383: 121125.
Mulvihill, M., A. Tao, K. Benjauthrit, J. Arnold and P. Yang. 2008. Surface‐enhanced Raman spectroscopy for trace
arsenic detection in contaminated water. Angewandte Chemie International Edition. 47(34): 6456–60.
Nassar, N. N. 2012. Iron oxide nanoadsorbents for removal of various pollutants from wastewater: an overview.
pp. 81–118. In: A. Bhatnagar [Ed.]. Application of Adsorbents for Water Pollution Control. Bentham Science
Publishers, India.
Niu, P., C. Fernández-Sánchez, M. Gich, C. Ayora and A. Roig. 2015. Electroanalytical assessment of heavy metals in
waters with bismuth nanoparticle-porous carbon paste electrodes. Electrochimica Acta. 165: 155–161.
Patil, P. O., P. V. Bhandari, P. K. Deshmukh, S. S. Mahale, A. G. Patil, H. R. Bafna, K. V. Patel and S. B. Bari. 2017.
Green fabrication of graphene-based silver nanocomposites using agro-waste for sensing of heavy metals. Res.
Chem. Intermed. 43(7): 3757–3773.
Pereira, F. A., K. S. Sousa, G. R. Cavalcanti, M. G. Fonseca, A. G. de Souza and A. P. Alves. 2013. Chitosan
montmorillonite biocomposite as an adsorbent for copper (II) cations from aqueous solutions. Int. J. Biol.
Macromol. 61: 471–478.
Petty, J. T., J. Zheng, N. V. Hud and R. M. Dickson. 2004. DNA-templated Ag nanocluster formation. J. Am. Chem.
Soc. 126(16): 5207–5212.
Praveen, A., E. Khan, D. S. Ngiimei, M. Perwez, M. Sardar and M. Gupta. 2018. Iron oxide nanoparticles as nano
adsorbents: a possible way to reduce arsenic phytotoxicity in Indian mustard plant (Brassica juncea L.). J.
Plant Growth Regul. 37(2): 612–624.
Rahman, A. F. A., A. A. Jalil, S. Triwahyono, A. Ripin, F. F. A. Aziz, N. A. A. Fatah, N. F. Jaafar, C. N. C. Hitam,
N. F. M. Salleh and N. S. Hassan. 2011. Strategies for introducing titania onto mesostructured silica
nanoparticles targeting enhanced photocatalytic activity of visible-light-responsive Ti-MSN catalysts. J. Clean
Prod. 143: 1–12.
Rotaru, A., S. Dutta, E. Jentzsch, K. Gothelf and A. Mokhir 2010. Selective dsDNA‐templated formation of copper
nanoparticles in solution. Angewandte Chemie International Edition. 49(33): 5665–7.
Sabzehei, K., S. H. Hadavi, M. G. Bajestani and S. Sheibani. 2020. Comparative evaluation of copper oxide nano
photocatalyst characteristics by formation of composite with TiO2 and Zno. Solid State Sci. 10: 106362.
Sakulthaew, C., C. Chokejaroenrat, A. Poapolathep, T. Satapanajaru and S. Poapolathep. 2017. Hexavalent chromium
adsorption from aqueous solution using carbon nano-onions (CNOs). Chemosphere. 184: 1168–1174.
Sauer, M. 2003. Single‐molecule‐sensitive fluorescent sensors based on photoinduced intramolecular charge transfer.
Angewandte Chemie International Edition. 42(16): 1790–3.
Schopf, C., A. Martín and D. Iacopino. 2017. Plasmonic detection of mercury via amalgam formation on surface-
immobilized single Au nanorods. Sci. Technol. Adv. Mater. 18(1): 60–67.
Remediation for Heavy Metal Contamination 291
Sha, J. C., C. Tong, H. Zhang, L. Feng and B. Liu. 2015. CdTe QDs functionalized mesoporous silica nanoparticles
loaded with conjugated polymers: a facile sensing platform for cupric (II) ion detection in water through
FRET. Dyes Pigments. 113: 102–9.
Shafi, A., S. Bano, N. Khan, S. Sultana, Z. Rehman, M. M. Rahman, M. M. Sabir, F. Coulon and M. Z. Khan. 2021.
Nanoremediation technologies for sustainable remediation of contaminated environments: recent advances
and challenges. Chemosphere. 130065.
Sharma, V., A. K. Saini and S. M. Mobin. 2016. Multicolour fluorescent carbon nanoparticle probes for live cell
imaging and dual palladium and mercury sensors. J. Mater. Chem. B. 4(14): 2466–2476.
Shawky, H. A. 2011. Improvement of water quality using alginate/montmorillonite composite beads. J. Appl. Polymer
Sci. 119(4): 2371–2378.
Sheet, I., A. Kabbani and H. Holail. 2014. Removal of heavy metals using nanostructured graphite oxide, silica
nanoparticles and silica/graphite oxide composite. Energy procedia. 50: 130–138.
Shirdel, B. and M. A. Behnajady. 2020. Visible-light induced degradation of rhodamine B by Ba-doped ZnO
Nanoparticles. J. Mol. Liq. 315: 113633.
Siddiqui, M. H. and M. H. Al-Whaibi. 2014. Role of nano-SiO2 in germination of tomato (Lycopersicum esculentum
seeds Mill.). Saudi J. Biol. Sci. 21: 13–17.
Stietiya, M. H. and J. J. Wang. 2014. Zinc and cadmium adsorption to aluminum oxide nanoparticles affected by
naturally occurring ligands. J. Environ. Qual. 43(2): 498–506.
Sundaram, R. M., A. Sekiguchi, M. Sekiya, T. Yamada and K. Hata. 2018. Copper/carbon nanotube composites:
research trends and outlook. Royal Soc. Open Sci. 5(11): 180814.
Tabish, T. A., F. A. Memon, D. E. Gomez, D. W. Horsell and S. Zhang. 2018. A facile synthesis of porous graphene
for efficient water and wastewater treatment. Sci. Rep. 8(1): 1–14.
Tang, W., Y. Su, Q. Li, S. Gao and J. K. Shang. 2013. Superparamagnetic magnesium ferrite nanoadsorbent for
effective arsenic (III, V) removal and easy magnetic separation. Water Res. 47(11): 3624–3634.
Ting, S. L., S. J. Ee, A. Ananthanarayanan, K. C. Leong and P. Chen. 2015. Graphene quantum dots functionalized
gold nanoparticles for sensitive electrochemical detection of heavy metal ions. Electrochimica Acta.
172: 7–11.
Tresintsi, S., K. Simeonidis, S. Estradé, C. Martinez-Boubeta, G. Vourlias, F. Pinakidou, M. Katsikini, E. C. Paloura,
G. Stavropoulos and M. Mitrakas. 2013. Tetravalent manganese feroxyhyte: a novel nanoadsorbent equally
selective for As (III) and As (V) removal from drinking water. Environ. Sci. Technol. 47(17): 9699–9705.
Varun, S. and S. C. Daniel. 2018. Emerging nanosensing strategies for heavy metal detection. Nanotechnol. Sustain.
Water Resour. 5: 199–255.
Xie, Y. L., S. Q. Zhao, H. L. Ye, J. Yuan, P. Song and S. Q. Hu. 2015. Graphene/CeO2 hybrid materials for the
simultaneous electrochemical detection of cadmium (II), lead (II), copper (II), and mercury (II). J. Electroanal.
Chem. 15: 757: 235–242.
Yaqoob, S., F. Ullah, S. Mehmood, T. Mahmood, M. Ullah, A. Khattak and M. A. Zeb. 2018. Effect of waste water
treated with TiO2 nanoparticles on early seedling growth of Zea mays L. J. Water Reuse Desalin. 8(3):
424–431.
Zhang, C., J. Sui, Y. Li, Tang and W. Cai. 2012. Efficient removal of heavy metal ions by thiol-functionalized
superparamagnetic carbon nanotubes. Chem. Eng. J. 210: 45–52.
Zhang, J., L. He, P. Chen, C. Tian, J. Wang, B. Liu, C. Jiang and Z. Zhang. 2017. A silica-based SERS chip for rapid
and ultrasensitive detection of fluoride ions triggered by a cyclic boronate ester cleavage reaction. Nanoscale.
9(4): 1599–606.
Zhang, R. and W. Chen. 2014. Nitrogen-doped carbon quantum dots: facile synthesis and application as a “turn-off”
fluorescent probe for detection of Hg2+ ions. Biosens. Bioelectron. 55: 83–90.
Zhao, L., W. Gu, C. Zhang, X. Shi and Y. Xian. 2016. In situ regulation nanoarchitecture of Au nanoparticles/reduced
graphene oxide colloid for sensitive and selective SERS detection of lead ions. J. Colloid Interface Sci. 465:
279–285.
Chapter 17
Modification Strategies
of g-C3N4 for Potential
Applications in Photocatalysis
A Sustainable Approach towards the
Environment
Sachin Shoran,1 Sweety Dahiya,1 Anshu Sharma2
and Sudesh Chaudhary 1,*
17.1 Introduction
Graphitic carbon nitride (g-C3N4) has been used as a visible-light-active polymeric photocatalyst
(460 nm) with a bandgap of 2.7 eV. g-C3N4 has become useful in engineering and physics because
of its low-cost and eco-friendly synthesis techniques. It is also a stable catalyst with promising
physicochemical properties (Ong et al. 2016). In comparison with other semiconductors, g-C3N4 can
be successfully prepared by multiple processes and has the desired electrical structures and
morphologies, as well as good thermal stability in the air up to 600ºC (Liu et al. 2021, Taha et al.
2021). The most frequently used precursors of g-C3N4 are urea, thiourea, dicyandiamide, cyanamide,
melamine and ammonium thiocyanate (Figure 17.2). Among several forms of carbon nitrides, such
as cubic C3N4, pseudocubic C3N4, α-C3N4 and β-C3N4; g-C3N4 is the most stable phase under
ambient circumstances (Ong et al. 2016). Researchers have proposed many strategies for improving
the performance and modulating the characteristics of g-C3N4, such as heterojunction with other
materials and doping with metal sulfides, noble metals, nonmetals and metal oxides nanoparticles
are examples of these materials (Ren et al. 2019, Jiang et al. 2020, Wei et al. 2022). Metal oxides
enhance the efficiency of g-C3N4, such as boosting light absorption and lowering the electron and
hole’s recombination by encouraging charge carrier separation. This is primarily because of their
appropriate band structure (Li et al. 2016, Fu et al. 2018). The g-C3N4 structure has been extensively
1
Center of Excellence for Energy and Environmental Studies, Deenbandhu Chhotu Ram University of Science and
Technology, Murthal-131039 (Haryana), India.
2
Department of Physics under School of Engineering and Technology, Central University of Haryana, Mahendergarh-123031
(Haryana), India.
* Corresponding author: [email protected]
Modification Strategies of g-C3N4 for Potential Applications in Photocatalysis 293
employed in several applications, particularly those connected to energy. The energy needed to make
electricity and heat will double by 2050, primarily because of industrialization, urbanization and
population growth (Dai et al. 2012) ). Oil, coal and other fossil fuels should be used less often since
they negatively influence the environment (Ong et al. 2016). Solar power and photocatalysis are two
solutions (Hasanvandian et al. 2022). Both need appropriate semiconductors, such as g-C3N4, that
have high activity for various catalytic processes, including water splitting, hydrogen production,
the degradation of organic pollutants and the conversion of CO2 (Zou et al. 2018, Arumugam
et al. 2022, Yang et al. 2022). Additionally, g-C3N4 can be utilized to clean wastewater and control
microorganisms (Zhang et al. 2019).
This chapter has focused on several broad details on the characterization and structure of bare
g-C3N4. Some possible adjustments, including doping to improve the characteristics of g-C3N4 are
also discussed. The studies on altering the structure and characteristics of g-C3N4 to increase its
effectiveness for various applications by mixing it with metals, metal oxides and nonmetals have
been described. In the final section of this chapter, the authors have made several recommendations
for prospective future studies in this area which, to our knowledge, have not yet been done.
Figure 17.2(a). Synthesis precursors for g-C3N4 preparation; (b) synthesis procedure using urea, thiourea, cyanamide and
melamine for g-C3N4.
Figure. 17.2. (a) Synthesis precursors for g-C3N4 preparation; (b) synthesis procedure using cyanamide, urea,
electron pairs on pz orbitals. This results in the formation of a highly delocalized conjugated system
thiourea, and melamine for g-C3N4.
(Hao et al. 2020, Li et al. 2019). The electron-hole recombination rate of bulk g-C3N4 is relatively
high because of the presence of low-coordinated N atoms in both the Conduction band (CB) and the
valence band (VB) that forms low delocalization bonds.
Most researchers are interested in how to modify g-C3N4 to improve its photocatalytic
performance (Mishra et al. 2019, Zhang et al. 2019), such as converting energy, breaking down
pollutants (Ren et al. 2019) and controlling microbes, etc. (Zhang et al. 2019).
Even though some studies offer suggestions for how to alter g-C3N4, a comprehensive analysis of
the fabrication of g-C3N4 with varied dimensionalities and their implications on diverse applications
connected to the environment and energy has not yet been done. This chapter provides an up-to-date
review of how to make g-C3N4 with different dimensions and how it can be used for energy and the
environment. Additionally, an overview of the research’s state and suggestions on how g-C3N4 will
develop in the future and what it could be used for has been provided.
Figure 17.3. Modification methods for g-C3N4 and various doping strategies.
Figure. 17.3. Modification methods for g-C3N4 and various doping strategies.
the ITO substrate was placed above the thiourea and melamine mixture in a crucible before it was
moved to the muffle furnace. There are some benefits and drawbacks of CVD. CVD does not require
a high vacuum to manufacture high-purity composites and may deposit various materials.
In contrast, a high vacuum environment is needed for Physical Vapor Deposition (PVD), which
includes sputtering. The disadvantage of CVD is that several CVD precursors, such as Ni (CO)4,
B2H6 and SiCl4, are expensive, extremely poisonous, explosive or corrosive. This process can also
produce dangerous byproducts such as CO or HF. The substrates are restricted since they need to
withstand high temperatures.
a solution of dicyandiamide. With this method, the properties of g-C3N4 can be changed by using
different precursors in different solvents. Li et al. (2004) were able to show that highly crystalline
g-C3N4 thin films could be made at room temperature using cyanuric chloride and an acetonitrile
melamine solution. This was done by a simple electrochemical deposition. The shape and size of
g-C3N4 particles were controlled by a combination of the templating method and electrochemical
deposition. Bai et al. (2010) reported hollow g-C3N4 microspheres were produced using different-
sized silica nanospheres. The spheres have an average diameter of about 1 m and are made up of
nanoparticles that range in size from 5 to 30 nm.
Figure 17.4. Synthetic approaches of g-C3N4 including (a) solid-state reaction, (b) electrochemical deposition,
Figure. 17.4. Synthetic approaches of g-C3N4 including (a) solid-state reaction, (b) electrochemical
(c) solvothermal reaction and (d) thermal decomposition.
deposition, (c) solvothermal reaction and (d) thermal decomposition.
band gaps and improved surface area for efficient charge transfer. The adsorption efficiency of Na
doped g-C3N4 produced from cyanamide and NaCl as precursors were reported by (Fronczak et al.
2017). Textural characteristics of Na-doped g-C3N4 were studied using N2-adsorption/desorption
isotherms. The findings showed that Na concentration was increased. The photocatalytic activity
and visible light absorption of g-C3N4 were significantly enhanced by deprotonation with Na+.
Similarly (Hu et al. 2014), created potassium (K) doped g-C3N4 with a band gap optimized
for removing RhB dye under visible light irradiation using dicyandiamide (DCDA) monomer and
potassium hydrate as precursors. Due to its enhanced electronic structure and redox potentials for
adequate consumption of photogenerated electrons, it was inferred from several investigations that
Na-doped g-C3N4 demonstrated higher photocatalytic capacity. Strong interactions between metal
dopants and lone pair electrons on g-C3N4 nitrogen pots make it easy for metal cations to get into
the framework. Alkali metal addition showed stable chemical activity despite being very reactive,
which justifies the conclusion that it will increase the photodegradation of organic contaminants.
the C–F bonding, which reduced the band gap of g–C3N4 from 2.69 eV to 2.63 eV. Additionally, DFT
studies showed that adding F to the carbon in the bay pushes the VB and CB to higher energy values.
F-doped g-C3N4 displayed around 2.7 folds more activity than untreated g-C3N4 in photocatalytic
hydrogen evolution.
For the fabrication of S doped g-C3N4 photocatalyst (Wang et al. 2015), used melamine and
thiourea precursors calcined at 520°C. The findings of photocatalytic reduction of CO2 showed that
the CH3OH yield with pure g-C3N4 was 0.81 mol g–1, whereas it was 1.12 mol g–1 for g-C3N4 doped
with S. Using melamine and thiourea as common precursors (Liang et al. 2016), created a series of
S doped g-C3N4 grafted with zinc phthalocyanines (ZnTNPc). S-doped g-C3N4 and ZnTNPc showed
a synergistic relationship for the photocatalytic elimination of Methylene blue dye, which was
4.5-fold more than that of zinc phthalocyanines (ZnTNPc). According to Mott-Schottky Relationship,
adding S atoms narrows the band gap and lowers the Conduction Band (CB) from 1.04V to 0.83V,
which facilitates the photocatalytic activity of MB.
By condensing oxalic acid and urea at a high temperature of 550°C (Qiu et al. 2017), created
porous O-doped g-C3N4. The band gap was reduced from 2.91 eV to 2.07 eV due to the inclusion of
the O atom into the g-C3N4 lattice. To enable the oxidation of benzene to phenol and other non-toxic
chemicals under the influence of visible light, a fluorinated g-C3N4 heterogeneous photocatalyst was
produced. The preparation was initiated by adding NH4F by thermal precursor in a consistent, easy
one-pot facile thermal polymerization technique to design and create B-doped g-C3N4 nanosheets.
Table 17.2. A list of CO2 reduction applications of the metal oxide-based g-C3N4 photocatalysts.
Table 17.3. List of the studies carried out on organic pollutant degradation by g-C3N4-based heterojunctions.
to g-C3N4-ZnO heterojunction, has a strong capacity in the degradation of RhB (Li et al. 2019,
Xia et al. 2019). Table 17.3 lists many ongoing research projects on the photodegradation applications
of g-C3N4-based heterojunctions.
17.5.2 Sensors
A g-C3N4 nanosheet is an excellent option for a modified electrode for sensors that can detect
analytes like dopamine, hydrogen peroxide, glucose, etc. These advantages include outstanding
fluorescence quenching abilities, quick response to external stimulations, high sensitivity to
analytes, high level of stability light and electricity conversion properties and biocompatibility
(Zou et al. 2018, Wang et al. 2019). As gas sensors, metal oxide semiconductor/g-C3N4 composites
are frequently utilized (Rahman et al. 2021). Consequently, the metal oxide-loaded g-C3N4 has also
disclosed new sensors to identify various materials. g-C3N4-TiO2-based structures are one of the
most popular composites for sensing and other applications. Due to this, the composite exhibited
exceptional stability, repeatability and excellent selectivity, another heterojunction utilized in UV-
assisted gas sensors is ZnO-g-C3N4. It is demonstrated that ZnO-g-C3N4 has significantly greater
ethanol (C2H5OH) detecting capacity than bare ZnO and g-C3N4. The best sensing performance
was demonstrated by the ZnO containing 8% g-C3N4, which is attributed to the efficient separation
of electrons and holes between g-C3N4 and ZnO and the catalytic impact of UV light at room
temperature (Zhai et al. 2018). The applications of g-C3N4-metal oxide heterojunctions as a sensor
material are listed in Table 17.4.
304 Bioremediation for Sustainable Environmental Cleanup
Table 17.4. The list of sensing-related applications for g-C3N4-metal oxide heterojunction.
17.6 Conclusion
g-C3N4 is a metal-free semiconductor with an adjustable band gap, excellent chemical and thermal
stability and appealing electrical characteristics. Graphitic carbon nitride absorbs UV light and has
a band gap of 2.7 eV. Modifying g-C3N4 via doping and blending with other materials to produce
composites can result in improved optoelectronic characteristics, and the composites can exhibit
synergistic properties. Many studies used doping elements to increase the effectiveness of bare
g-C3N4 light harvesting, while mixing g-C3N4 with other materials, such as metals, metal oxides and
nonmetals is another strategy to deal with this issue. The examination of g-C3N4 heterojunctions with
designed bandgaps and optimized surfaces to improve the absorption spectrum towards the visible-
light region, reduced charge carrier recombination, and increased surface adsorption and reaction
are among the critical areas of g-C3N4 based material research. The highlighted applications of
modified g-C3N4 included water splitting, bacterial disinfection, energy storage, photodegradation
of organic pollutants, CO2 reduction, sensing, etc. Even though most recent research has shown that
g-C3N4-based photocatalysts work very well, their full potential has not yet been fully realized. The
main problems are in finding a green way to make a photocatalyst with a high surface area and good
photostability, testing how well a g-C3N4 based photocatalyst works with real industrial wastewater
and improving reactor design to get the best photocatalytic activity.
References
Acharya, R. and P. Kulamani. 2020. A review on TiO2/g-C3N4 visible-light-responsive photocatalysts for sustainable
energy generation and environmental remediation. J. Environ. Chem. Eng. 8(4).
Adhikari, S. and Do H. Kim. 2020. Heterojunction C3N4/MoO3 microcomposite for highly efficient photocatalytic
oxidation of Rhodamine B. Appl. Surf. Sci. 511.
Ahmad, R., N. Tripathy, A. Khosla, M. Khan, P. Mishra, W. A. Ansari, M. A. Syed and Y.-B. Hahn. 2020. Review—
recent advances in nanostructured graphitic carbon nitride as a sensing material for heavy metal ions. J.
Electrochem. Soc. 167(3): 037519.
306 Bioremediation for Sustainable Environmental Cleanup
Gu, Y., A. Bao, X. Zhang, J. Yan, Q. Du, M. Zhang and X. Qi. 2021. Facile fabrication of sulfur-doped Cu2O and
g-C3N4 with Z-Scheme structure for enhanced photocatalytic water splitting performance. Mater. Chem. Phys.
266: 124542.
Guo, H., S. Wan, Y. Wang, W. Ma, Q. Zhong and J. Ding. 2021. Enhanced photocatalytic CO2 reduction over direct
Z-Scheme NiTiO3/g-C3N4 nanocomposite promoted by efficient interfacial charge transfer. Chem. Eng. J.
412: 12864.
Guo, Q., Y. Xie, X. Wang, S. Lv, T. Hou and X. Liu. 2003. Characterization of well-crystallized graphitic carbon
nitride nanocrystallites via a benzene-thermal route at low temperatures. Chem. Phys. Lett. 380(1-2): 84–87.
Guo, Y., B. Chang, T. Wen, S. Zhang, M. Zeng, N. Hu, Y. Su, Z. Yang and B. Yang. 2020. A Z-scheme photocatalyst
for enhanced photocatalytic H2 evolution, constructed by growth of 2D plasmonic MoO3-x nanoplates onto
2D g-C3N4 nanosheets. J. Colloid Interface Sci. 567: 213–223.
Han, C., L. Ge, C. Chen, Y. Li, X. Xiao, Y. Zhang and L. Guo. 2014. Novel visible light induced Co3O4-g-C3N4
heterojunction photocatalysts for efficient degradation of methyl orange. Appl. Catal. B: Environ. 147:
546–553.
Hao, J., S. Zhang, F. Ren, Z. Wang, J. Lei, X. Wang, T. C. and L. Li. 2017. Synthesis of TiO2@g-C3N4 core-shell
nanorod arrays with Z-scheme enhanced photocatalytic activity under visible light. J. Colloid Interface Sci.
508: 419–425.
Hao, Q., C. Xie, Y. Huang, D. Chen, Y. Liu, W. Wei and B.-J. Ni. 2020. Accelerated Separation of Photogenerated
Charge Carriers and Enhanced Photocatalytic Performance of g-C 3: 249–258.
Hasanvandian, F., M. Moradi, S. A. Samani, B. Kakavandi, S. R. Setayesh and M. Noorisepehr. 2022. Effective
promotion of g–C3N4 photocatalytic performance via surface oxygen vacancy and coupling with bismuth-
based semiconductors towards antibiotics degradation. Chemosphere. 287 (Pt 3).
Hieu, V. Q., T. C. Lam, A. Khan, T. T. T. Vo, T. Q. Nguyen, V. D. Doan, D. L. Tran, V. T. Le and V. A. Tran. 2021.
TiO2/Ti3C2/g-C3N4 ternary heterojunction for photocatalytic hydrogen evolution. Chemosphere. 285: 13142.
Hu, S., F. Li, Z. Fan, F. Wang, Y. Zhao and Z. Lv. 2014. Band gap-tunable potassium doped graphitic carbon nitride
with enhanced mineralization ability. Dalton Transact. 44(3): 1084–1092.
Hu, S., L. Ma, J. You, F. Li, Z. Fan, F. Wang, D. Liu and J. Gui. 2014. A simple and efficient method to prepare a
phosphorus modified G-C 3N4 visible light photocatalyst. RSC Adv. 4(41): 21657–21663.
Huan, Z., J. Chang and J. Zhou. 2010. Low-temperature fabrication of macroporous scaffolds through foaming and
hydration of tricalcium silicate paste and their bioactivity. J. Mater. Sci. 45(4): 961–968.
Ji, C., S. N. Yin, S. Sun and S. Yang. 2018. An in situ mediator-free route to fabricate Cu2O/g-C3N4 Type-II
heterojunctions for enhanced visible-light photocatalytic H2 generation. Appl. Surf. Sci. 434: 1224–1231.
Jiang, H., Y. Li, D. Wang, X. Hong and B. Liang. 2020. Recent advances in heteroatom doped graphitic carbon nitride
(g-C3N4) and g-C3N4/Metal oxide composite photocatalysts. Curr. Organ. Chem. 24(6): 673–693.
Jiang, L., X. Yuan, Y. Pan, J. Liang, G. Zeng, Z. Wu and H. Wang. 2017. Doping of graphitic carbon nitride for
photocatalysis: a reveiw. Appl. Catal. B: Environ. 217: 388–406.
Jiang, Y., Z. Sun, Q. Chen, C. Cao, Y. Zhao, W. Yang, L. Zeng and L. Huang. 2022. Fabrication of 0D/2D TiO2
Nanodots/g-C3N4 S-Scheme heterojunction photocatalyst for efficient photocatalytic overall water splitting.
Appl. Surf. Sci. 571: 15128.
Jin, C., M. Wang, Z. Li, J. Kang, Y. Zhao, J. Han and Z. Wu. 2020. Two dimensional Co3O4/g-C3N4 Z-scheme
heterojunction: mechanism insight into enhanced peroxymonosulfate-mediated visible light photocatalytic
performance. Chem. Eng. J. 398: 125569.
Jin, R., S. Hu, J. Gui and D. Liu. 2015. A convenient method to prepare novel rare earth metal ce-doped carbon nitride
with enhanced photocatalytic activity under visible light. Bull. Korean Chem. Soc. 36(1): 17–23.
Karimi, M. A., M. Atashkadi, M. Ranjbar and A. Habibi-Yangjeh. 2020. Novel visible-light-driven photocatalyst of
NiO/Cd/g-C3N4 for enhanced degradation of methylene blue. Arab. J. Chem. 13(6): 5810–5820.
Karpuraranjith, M., Y. Chen, S. Rajaboopathi, M. Ramadoss, K. Srinivas, D. Yang and B. Wang. 2022. Three-
dimensional porous MoS2 nanobox embedded g-C3N4@TiO2 architecture for highly efficient photocatalytic
degradation of organic pollutant. J. Colloid Interface Sci. 605: 613–623.
Kočí, K., M. Reli, I. Troppová, M. Šihor, J. Kupková, P. Kustrowski and P. Praus. 2017. Photocatalytic decomposition
of N2O over TiO2/g-C3N4 Photocatalysts Heterojunction. Appl. Surf. Sci. 396: 1685–1695.
Kong, Y., C. Lv, C. Zhang and G. Chen. 2020. Cyano group modified G-C3N4: molten salt method achievement and
promoted photocatalytic nitrogen fixation activity. Appl. Surf. Sci. 515: 14600.
Kuriki, R., K. Sekizawa, O. Ishitani and K. Maeda. 2015. Visible-light-driven CO2 reduction with carbon nitride:
enhancing the activity of ruthenium catalysts. In Angewandte Chemie - Int. Ed. 54: 2406–2409. 1002/
anie.201411170: 54.
Lan, Z. A., G. Zhang and X. Wang. 2016. A facile synthesis of Br-modified g-C3N4 semiconductors for photoredox
water splitting. Appl. Catal. B: Environ. 192: 116–125.
308 Bioremediation for Sustainable Environmental Cleanup
Li, C., C. B. Cao and H. S. Zhu. 2004. Graphitic carbon nitride thin films deposited by electrodeposition. Mater. Lett.
58(12-13): 1903–1906.
Li, G., X. Nie, J. Chen, Q. Jiang, T. An, P. K. Wong, H. Zhang, H. Zhao and H. Yamashita. 2015. Enhanced
visible-light-driven photocatalytic inactivation of Escherichia coli using g-C3N4/TiO2 hybrid photocatalyst
synthesized using a hydrothermal-calcination approach. Water Res. 86: 17–24.
Li, J., C. Cao and H. Zhu. 2007. Synthesis and characterization of graphite-like carbon nitride nanobelts and
nanotubes. Nanotechnol. 18(11): 5.
Li, J., Y. Huan and Z. Zhu. 2016. Improved photoelectrochemical performance of Z-scheme g-C3N4/Bi2O3/BiPO4
heterostructure and degradation property. Appl. Surf. Sci. 385: 34–41.
Li, J., X. Zhang, F. Raziq, J. Wang, C. Liu, Y. Liu et al. 2017. Improved photocatalytic activities of G-C3N4 nanosheets
by effectively trapping holes with halogen-induced surface polarization and 2,4-dichlorophenol decomposition
mechanism. Appl. Catal. B: Environ. 218: 60–67.
Li, N., Y. Tian, J. Zhao, J. Zhang, W. Zuo, L. Kong and H. Cui. 2018. Z-scheme 2D/3D g-C3N4@ZnO with enhanced
photocatalytic activity for cephalexin oxidation under solar light. Chem. Eng. J. 352: 412–422.
Li, X., Y. Feng, M. Li, W. Li, H. Wei and D. Song. 2015. Smart hybrids of Zn2GeO4 nanoparticles and ultrathin
G-C3N4 layers: synergistic lithium storage and excellent electrochemical performance. Adv. Funct. Mater.
25(44): 6858–6866.
Li, X., J. Xiong, Y. Xu, Z. Feng and J. Huang. 2019. Defect-assisted surface modification enhances the visible light
photocatalytic performance of g-C3N4@C-TiO2 direct Z-scheme heterojunctions. Cuihua Xuebao/Chinese J.
Catal. 40(3): 424–433.
Li, Y., Z. Ruan, Y. He, J. Li, K. Li, Y. Yang, D. Xia, K. Lin and Y. Yuan. 2019. Enhanced photocatalytic H2 evolution
and phenol degradation over sulfur doped meso/macroporous g-C3N4 spheres with continuous channels. Int.
J. Hydrogen Energy. 44(2): 707–719.
Li, X., H. Jiang, C. Ma, Z. Zhu, X. Song, H. Wang, P. Huo and X. Li. 2021. Local surface plasma resonance effect
enhanced Z-scheme ZnO/Au/g-C3N4 film photocatalyst for reduction of CO2 to CO. Appl. Catal. B: Environ.
283: 11963.
Li, Y., S. Zhu, X. Kong, Y. Liang, Z. Li, S. Wu, C. Chang, S. Luo and Z. Cui. 2021. In situ synthesis of a novel
Mn3O4/g-C3N4 p-n heterostructure photocatalyst for water splitting. J. Colloid Int. Sci. 586: 778–784.
Liang, Q., M. Zhang, C. Liu, S. Xu and Z. Li. 2016. Sulfur-doped graphitic carbon nitride decorated with zinc
phthalocyanines towards highly stable and efficient photocatalysis. Appl. Catal. A: General. 519: 107–115.
Liebig, J. 1834. About some nitrogen compounds. Ann. Pharm. 10: 10.
Liu, A. Y. and M. L. Cohen. 1989. Prediction of new low compressibility solids. Sci. 245(4920): 841–842.
Liu, L., Xi L., Y. Li, F. Xu, Z. Gao, X. Zhang, Y. Song, H. Xu and H. Li. 2018. Facile synthesis of few-layer g-C3N4/
ZnO composite photocatalyst for enhancing visible light photocatalytic performance of pollutants removal.
Colloids Surf. A: Physicochem. Eng. Asp. 537: 516–523.
Liu, X., R. Ma, L. Zhuang, B. Hu, J. Chen, X. Liu and X. Wang. 2021. Recent developments of doped G-C3N4
photocatalysts for the degradation of organic pollutants. Crit. Rev. Environ. Sci. Technol. 51(8): 751–790.
Liu, Y., J. Jiang, Y. Sun, S. Wu, Y. Cao, W. Gong and J. Zou. 2017. NiO and Co3O4 Co-Doped g-C3N4 nanocomposites
with excellent photoelectrochemical properties under visible light for detection of tetrabromobisphenol-A.
RSC Adv. 7(57): 36015–36020.
Liu, Y., H. Liu, H. Zhou, T. Li and L. Zhang. 2019. A Z-scheme mechanism of N-ZnO/g-C3N4 for enhanced H2
evolution and photocatalytic degradation. Appl. Surf. Sci. 466: 133–140.
Lu, X., L. Gai, D. Cui, Q. Wang, X. Zhao and X. Tao. 2007. Synthesis and characterization of C3N4 nanowires and
Pseudocubic C3N4 polycrystalline nanoparticles. Mater. Lett. 61(21): 4255–4258.
Luo, J., X. Zhou, L. Ma and X. Xu. 2015. Enhancing visible-light photocatalytic activity of g-C3N4 by doping
phosphorus and coupling with CeO2 for the degradation of methyl orange under visible light irradiation. RSC
Adv. 5(84): 68728–68735.
Ma, R., J. Guo, D. Wang, M. He, S. Xun, J. Gu, W. Zhu and H. Li. 2019. Preparation of highly dispersed WO3/
few layer g-C3N4 and its enhancement of catalytic oxidative desulfurization activity. Colloids Surf. A:
Physicochem. Eng. Asp. 572: 250–258.
Mao, L., X. Xue, X. Xu, W. Wen, M. M. Chen, X. Zhang and S. Wang. 2021. Heterostructured CuO-g-C3N4
nanocomposites as a highly efficient photocathode for photoelectrochemical Aflatoxin B1 sensing. Sens.
Actuators B: Chem. 329: 12914.
Marchal, C., T. Cottineau, M. G. Méndez-Medrano, C. Colbeau-Justin, V. Caps and V. Keller. 2018. Au/TiO2–GC3N4
nanocomposites for enhanced photocatalytic H2 production from water under visible light irradiation with
very low quantities of sacrificial agents. Adv. Energy Mater. 8(14): 1702142.
Matsunaga, T., R. Tomoda, T. Nakajima and H. Wake. 1985. Photoelectrochemical sterilization of microbial cells by
semiconductor powders. FEMS Microbiol. Lett. 29 (1-2): 211–214.
Modification Strategies of g-C3N4 for Potential Applications in Photocatalysis 309
Meenakshisundaram, I., S. Kalimuthu, P. P. Gomathi and S. Karthikeyan. 2019. Facile green synthesis and
antimicrobial performance of Cu2O nanospheres decorated G-C3N4 nanocomposite. Mater. Res. Bull. 112:
331–335.
Mishra, A., A. Mehta, S. Basu, N. P. Shetti, K. R. Reddy and T. M. Aminabhavi. 2019. Graphitic carbon nitride
(g-C3N4)-based metal-free photocatalysts for water splitting: a review. Carbon. 149: 693–721.
Mohammadi, I., F. Zeraatpisheh, E. Ashiri and K. Abdi. 2020. Solvothermal synthesis of G-C3N4 and ZnO
nanoparticles on TiO2 nanotube as photoanode in DSSC. Int. J. Hydrogen Energy. 45(38): 18831–18839.
Montigaud, H., B. Tanguy, G. Demazeau, I. Alves and S. Courjault. 2000. Dream or Reality? Solvothermal Synthesis
as Macroscopic Samples of the C. 3: 2547–2552.
Mulik, B. B., B. D. Bankar, A. V. Munde, P. P. Chavan, A. V. Biradar and B. R. Sathe. 2021. Electrocatalytic and
catalytic CO2 hydrogenation on ZnO/g-C3N4 hybrid nanoelectrodes. Appl. Surf. Sci. 538: 14812.
Ngullie, R. C., O. A. Saleh, B. Kandasamy, P. Shanmugam, T. Pazhanivel and P. Arunachalam. 2020. Synthesis
and characterization of efficient ZnO/g-C3N4 nanocomposites photocatalyst for photocatalytic degradation of
methylene blue. Coatings 10(5): 500.
Nie, N., L. Zhang, J. Fu, B. Cheng and J. Yu. 2018. Self-assembled hierarchical direct Z-Scheme g-C3N4/ZnO
microspheres with enhanced photocatalytic CO2 reduction performance. Appl. Surf. Sci. 441: 12–22.
Ong, W. J., L. L. Tan, Y. H. Ng, S. T. Yong and S. P. Chai. 2016. Graphitic carbon nitride (g-C3N4)-based photocatalysts
for artificial photosynthesis and environmental remediation: are we a step closer to achieving sustainability?
Chem. Rev. 116(12): 7159–7329.
Ouyang, K., B. Xu, C. Yang, H. Wang, P. Zhan and S. Xie. 2022. Synthesis of a novel Z-scheme Ag/WO3/g
C3N4 nanophotocatalyst for degradation of oxytetracycline hydrochloride under visible light. Mater. Sci.
Semiconductor Process. 137.
Pan, T., D. Chen, W. Xu, J. Fang, S. Wu, Z. Liu, K. Wu and Z. Fang. 2020. Anionic polyacrylamide-assisted
construction of thin 2D-2D WO3/g-C3N4 step-scheme heterojunction for enhanced tetracycline degradation
under visible light irradiation. J. Hazard. Mater. 393: 12236.
Pant, B., M. Park, J. H. Lee, H. Y. Kim and S. J. Park. 2017. Novel magnetically separable silver-iron oxide
nanoparticles decorated graphitic carbon nitride nano-sheets: a multifunctional photocatalyst via one-step
hydrothermal process. J. Colloid Interface Sci. 496: 343–352.
Paul, A. M., A. Sajeev, R. Nivetha, K. Gothandapani, P. Bhardwaj, K. Govardhan et al. 2020. Cuprous oxide (Cu2O)/
Graphitic Carbon Nitride (g-C3N4) nanocomposites for electrocatalytic hydrogen evolution reaction. Diamond
Relat. Mater. 107: 107899.
Paul, D. R., S. Gautam, P. Panchal, S. P. Nehra, P. Choudhary and A. Sharma. 2020. ZnO-modified g-C3N4: a potential
photocatalyst for environmental application. ACS Omega. 5(8): 3828–3838.
Qiu, P., C. Xu, H. Chen, F. Jiang, X. Wang, R. Lu and X. Zhang. 2017. One step synthesis of oxygen doped porous
graphitic carbon nitride with remarkable improvement of photo-oxidation activity: role of oxygen on visible
light photocatalytic activity. Appl. Catal. B: Environ. 206(1016): 319–327.
Qu, A., X. Xu, H. Xie, Y. Zhang, Y. Li and J. Wang. 2016. Effects of calcining temperature on photocatalysis of
G-C3N4/TiO2 composites for hydrogen evolution from water. Mater. Res. Bull. 80: 167–176.
Rabani, I., R. Zafar, K. Subalakshmi, H. S. Kim, C. Bathula and Y. S. Seo. 2021. A facile mechanochemical preparation
of Co3O4@g-C3N4 for application in supercapacitors and degradation of pollutants in water. J. Hazard. Mater.
407.
Rahman, N., J. Yang, Zulfiqar, M. Sohail, R. Khan, A. Iqbal, C. Maouche et al. 2021. Insight into metallic oxide
semiconductor (SnO2, ZnO, CuO, α-Fe2O3, WO3)-Carbon Nitride (g-C3N4) heterojunction for gas sensing
application. Sens. Actuators A: Phys. 332: 11312.
Ramachandra, M., S. D. Kalathiparambil, R. Pai, J. R. Jaleel and D. Pinheiro. 2020. Improved photocatalytic activity
of G-C3N4/ZnO: A PotentialDirect Z-Scheme Nanocomposite. Chem. Sel. 5(38): 11986–11995.
Rathi, A. K., H. Kmentová, A. Naldoni, A. Goswami, M. B. Gawande, R. S. Varma, S. Kment and R. Zboril. 2018.
Significant Enhancement of photoactivity in hybrid TiO2/g-C3N4 nanorod catalysts modified with Cu-Ni
based nanostructures. ACS Appl. Nano Mater. 1(6): 2526–2535.
Ren, Y., D. Zeng and W. J. Ong. 2019. Interfacial engineering of graphitic Carbon Nitride (g-C3N4)-based metal
sulfide heterojunction photocatalysts for energy conversion: a review. Cuihua Xuebao/Chinese J. Catal. 40(3):
289–319.
Selvarajan, S., A. Suganthi and M. Rajarajan. 2018. Fabrication of G-C3N4/NiO heterostructured nanocomposite
modified glassy carbon electrode for quercetin biosensor. Ultrason. Sonochem. 41: 651–660.
Shaikh, A.V., R. S. Mane, O. S. Joo, S. H. Han and H. M. Pathan. 2017. Electrochemical deposition of cadmium
selenide films and their properties: a review. J. Solid State Electrochem. 21(9): 2517–2530.
Shan, W., Y. Hu, Z. Bai, M. Zheng and C. Wei. 2016. In situ preparation of G-C3N4/bismuth-based oxide
nanocomposites with enhanced photocatalytic activity. Appl. Catal. B: Environ. 188: 1–12.
310 Bioremediation for Sustainable Environmental Cleanup
Shandilya, P., D. Mittal, A. Sudhaik, M. Soni, P. Raizada, A. K. Saini and P. Singh. 2019. GdVO4 modified fluorine
doped graphene nanosheets as dispersed photocatalyst for mitigation of phenolic compounds in aqueous
environment and bacterial disinfection. Sep. Purif. Technol. 210: 804–816.
Shanmugam, V., K. S. Jeyaperumal, P. Mariappan and A. L. Muppudathi. 2020. Fabrication of novel G-C3N4 based
MoS2 and Bi2O3 nanorod embedded ternary nanocomposites for superior photocatalytic performance and
destruction of bacteria. New J. Chem. 44(30): 13182–13194.
Shi, J. W., Y. Zou, L. Cheng, D. Ma, D. Sun, S. Mao, L. Sun, C. He and Z. Wang. 2019. In-situ phosphating to
synthesize Ni2P decorated NiO/g-C3N4 p-n junction for enhanced photocatalytic hydrogen production. Chem.
Eng. J. 378.
Shi, J., B. Zheng, L. Mao, C. Cheng, Y. Hu, H. Wang, G. Li, D. Jing and X. Liang. 2021. MoO3/g-C3N4 Z-scheme
(S-Scheme) system derived from MoS2/melamine dual precursors for enhanced photocatalytic H2 evolution
driven by visible light. Int. J. Hydrogen Energy. 46(3): 2927–2935.
Shoran, S., S. Chaudhary and A. Sharma. 2022. Photocatalytic dye degradation and antibacterial activities of CeO2/g
C3N4 nanomaterials for environmental applications. Environ. Sci. Pollut. Res.
Si, Y., X. Zhang, T. Liang, X. Xu, L. Qiu, P. Li and S. Duo. 2020. Facile in-situ synthesis of 2D/3D g-C3N4/Cu2O
heterojunction for high-performance photocatalytic dye degradation. Mater. Res. Express. 7(1): 15524.
Song, T., X. Zhang and P. Yang. 2021. Bifunctional nitrogen-doped carbon dots in g-C3N4/WOx heterojunction for
enhanced photocatalytic water-splitting performance. Langmuir. 37(14): 4236–4247.
Sudrajat, H. 2018. A one-pot, solid-state route for realizing highly visible light active Na-Doped GC3N4 photocatalysts.
J. Solid State Chem. 257: 26–33.
Sumathi, M., A. Prakasam and P. M. Anbarasan. 2019. Fabrication of hexagonal disc shaped nanoparticles G-C3N4/
NiO heterostructured nanocomposites for efficient visible light photocatalytic performance. J. Cluster Sci.
3(30): 757–766.
Sun, Y., J. Jiang, Y. Liu, S. Wu and J. Zou. 2018. A facile one-pot preparation of Co3O4/g-C3N4 heterojunctions with
excellent electrocatalytic activity for the detection of environmental phenolic hormones. Appl. Surf. Sci. 430:
362–370.
Sun, Z., W. Fang, L. Zhao, H. Chen, X. He, W. Li, P. Tian and Z. Huang. 2019. G-C3N4 foam/Cu2O QDs with excellent
CO2 adsorption and synergistic catalytic effect for photocatalytic CO2 reduction. Environ. Int. 130: 10489.
Taha, M. M., L. G. Ghanem, M. A. Hamza and N. K. Allam. 2021. Highly stable supercapacitor devices based on
three-dimensional bioderived carbon encapsulated g-C3N4 nanosheets. ACS Appl. Energy Mater. 4(9): 10344–
10355.
Tan, X., X. Wang, H. Hang, D. Zhang, N. Zhang, Z. Xiao and H. Tao. 2019. Self-assembly method assisted synthesis
of g-C3N4/ZnO heterostructure nanocomposites with enhanced photocatalytic performance. Opt. Mater. 96.
Tang, J. Y., R. T. Guo, W. G. Zhou, C. Y. Huang and W. G. Pan. 2018. Ball-flower like NiO/g-C3N4 heterojunction for
efficient visible light photocatalytic CO2 reduction. Appl. Catal. B: Environ. 237: 802–810.
Tang, Z., C. Wang, W. He, Y. Wei, Z. Zhao and J. Liu. 2022. The Z-Scheme g-C3N4/3DOM-WO3 photocatalysts with
enhanced activity for CO2 photoreduction into CO. Chinese Chem. Lett. 33 (in press).
Teter, D. M. and R. J. Hemley. 1996. Low-compressibility carbon nitrides. Science. 271(5245): 53–55.
Thomas, J., K. S. Ambili and S. Radhika. 2018. Synthesis of Sm3+-doped graphitic carbon nitride nanosheets for the
photocatalytic degradation of organic pollutants under sunlight. Catal. Today. 310: 11–18.
Tonda, S., S. Kumar, S. Kandula and V. Shanker. 2014. Fe-doped and -mediated graphitic carbon nitride nanosheets
for enhanced photocatalytic performance under natural sunlight. J. Mater. Chem. A 2(19): 6772–6780.
Tran, H. H., P. H. Nguyen, M. L. P. Le, S.-J. Kim and V. Vo. 2019. SnO2 Nanosheets/Graphite Oxide/g-C3N4 composite
as enhanced performance anode material for lithium ion batteries. Chem. 715: 284–292.
Vignesh, S., S. Suganthi, J. K. Sundar and V. Raj. 2019. Construction of α-Fe2O3/CeO2 decorated g-C3N4 nanosheets
for magnetically separable efficient photocatalytic performance under visible light exposure and bacterial
disinfection. Appl. Surf. Sci. 488: 763–777.
Wang, H., H. Li, Z. Chen, J. Li, X. Li, P. Huo and Q. Wang. 2020. TiO2 modified G-C3N4 with enhanced photocatalytic
CO2 reduction performance. Solid State Sci. 100: 10609.
Wang, K., Q. Li, B. Liu, B. Cheng, W. Ho and J. Yu. 2015. Sulfur-doped g-C3N4 with enhanced photocatalytic CO2
Reduction Performance. Appl. Catal. B: Environ. 176-177: 44–52.
Wang, S., J. Long, T. Jiang, S. Li, D. Li, X. Xie and F. Xu. 2022. Magnetic Fe3O4/CeO2/g-C3N4 composites with
a visible-light response as a high efficiency fenton photocatalyst to synergistically degrade tetracycline. Sep.
Purif. Technol. 278: 11960.
Wang, W., Y. Zhan and G. Wang. 2001. One-step, solid-state reaction to the synthesis of copper oxide nanorods in the
presence of a suitable surfactant. Chem. Commun. 8(8): 727–728.
Wang, Y., R. Zhang, Z. Zhang, J. Cao and T. Ma. 2019. Host–guest recognition on 2D graphitic carbon nitride for
nanosensing. Adv Mater Interfaces. 6(23): 19014.
Modification Strategies of g-C3N4 for Potential Applications in Photocatalysis 311
Wei, T., J. Xu, C. Kan, L. Zhang and X. Zhu. 2022. Au tailored on G-C3N4/TiO2 heterostructure for enhanced
photocatalytic performance. J. Alloys Compd. 894.
Wen, P., Y. Sun, H. Li, Z. Liang, H. Wu, J. Zhang, H. Zeng, S. M. Geyer and L. Jiang. 2020. A highly active three-
dimensional Z-Scheme ZnO/Au/g-C3N4 photocathode for efficient photoelectrochemical water splitting. Appl.
Catal. B: Environ. 263.
Xia, Y., L. Xu, J. Peng, J. Han, S. Guo, L. Zhang, Z. Han and S. Komarneni. 2019. TiO2@g-C3N4 core/shell spheres
with uniform mesoporous structures for high performance visible-light photocatalytic application. Ceramics
Int. 45(15): 18844–18851.
Xu, D., X. Li, J. Liu and L. Huang. 2013. Synthesis and photocatalytic performance of europium-doped graphitic
carbon nitride. J. Rare Earths 31(11): 1085–1091.
Xu, J., Y. Li, X. Zhou, Y. Li, Z. D. Gao, Y. Y. Song and P. Schmuki. 2016. Graphitic C3N4-sensitized TiO2 nanotube
layers: a visible-light activated efficient metal-free antimicrobial platform. Chem. - A European J. 22(12):
3947–3951.
Yang, S., Q. Sun, W. Han, Y. Shen, Z. Ni, S. Zhang, L. Chen, L. Zhang, J. Cao and H. Zheng. 2022. A simple and
highly efficient composite based on G-C3N4for super rapid removal of multiple organic dyes from water under
sunlight. Catal. Sci. Technol. 12(3): 786–798.
Ye, L., D. Wang and S. Chen. 2016. Fabrication and enhanced photoelectrochemical performance of MoS2/S-doped
g-C3N4 heterojunction film. ACS Appl. Mater. Interfaces. 8(8): 5280–5289.
Zhai, J., T. Wang, C. Wang and D. Liu. 2018. UV-light-assisted ethanol sensing characteristics of g-C3N4/ZnO
composites at room temperature. Appl. Surface Sci. 441: 317–323.
Zhang, C., Y. Li, D. Shuai, Y. Shen, W. Xiong and L. Wang. 2019. Graphitic carbon nitride (g-C3N4)-based
photocatalysts for water disinfection and microbial control: a review. Chemosphere. 214: 462–479.
Zhang, M., X. Bai, D. Liu, J. Wang and Y. Zhu. 2015. Enhanced catalytic activity of potassium-doped graphitic carbon
nitride induced by lower valence position. Appl. Catal. B: Environ. 164: 77–81.
Zhang, P., X. Li, C. Shao and Y. Liu. 2015. Hydrothermal synthesis of carbon-rich graphitic carbon nitride nanosheets
for photoredox catalysis. J. Mater. Chem. A 3(7): 3281–3284.
Zhang, S., P. Gu, R. Ma, C. Luo, T. Wen, G. Zhao, W. Cheng and X. Wang. 2019. Recent developments in fabrication
and structure regulation of visible-light-driven g-C3N4-based photocatalysts towards water purification: a
critical review. Catal. Today. 335: 65–77.
Zhang, Y., J. Lu, M. R. Hoffmann, Q. Wang, Y. Cong, Q. Wang and H. Jin. 2015. Synthesis of G-C3N4/Bi2O3/TiO2
composite nanotubes: enhanced activity under visible light irradiation and improved photoelectrochemical
activity. RSC Adv. 5(60): 48983–48991.
Zhao, X., J. Guan, J. Li, X. Li, H. Wang, P. Huo and Y. Yan. 2021. CeO2/3D g-C3N4 heterojunction deposited with Pt
cocatalyst for enhanced photocatalytic CO2 reduction. Appl. Surf. Sci. 537.
Zhou, Y., L. Zhang, J. Liu, X. Fan, B. Wang, M. Wang, W. Ren, J. Wang, M. Li and J. Shi. 2015. Brand new P-doped
g-C3N4: enhanced photocatalytic activity for H2 evolution and Rhodamine B degradation under visible light.
J. Mater. Chem. A 3(7): 3862–3867.
Zhu, W., F. Sun, R. Goei and Y. Zhou. 2017. Construction of WO3-g-C3N4 composites as efficient photocatalysts for
pharmaceutical degradation under visible light. Catal. Sci. Technol. 7(12): 2591–2600.
Zhu, Y. P., T. Z. Ren and Z. Y. Yuan. 2015. Mesoporous phosphorus-doped g-C3N4 nanostructured flowers with
superior photocatalytic hydrogen evolution performance. ACS Appl. Mater. Interfaces. 7(30): 16850–16856.
Zou, J., S. Wu, Y. Liu, Y. Sun, Y. Cao, J. P. Hsu, A. T. S. Wee and J. Jiang. 2018. An ultra-sensitive electrochemical
sensor based on 2D g-C3N4/CuO nanocomposites for dopamine detection. Carbon. 130: 652–663.
Zou, W., B. Deng, X. Hu, Y. Zhou, Y. Pu, S. Yu, K. Ma, J. Sun, H. Wan and L. Dong. 2018. Crystal-plane-dependent
metal oxide-support interaction in CeO2/g-C3N4 for photocatalytic hydrogen evolution. Appl. Catal. B:
Environ. 238: 111–118.
Index
A F
B G
P S