Flocculation - Processes and Applications
Flocculation - Processes and Applications
FLOCCULATION
No part of this digital document may be reproduced, stored in a retrieval system or transmitted in any form or
by any means. The publisher has taken reasonable care in the preparation of this digital document, but makes no
expressed or implied warranty of any kind and assumes no responsibility for any errors or omissions. No
liability is assumed for incidental or consequential damages in connection with or arising out of information
contained herein. This digital document is sold with the clear understanding that the publisher is not engaged in
rendering legal, medical or any other professional services.
CHEMISTRY RESEARCH AND APPLICATIONS
FLOCCULATION
ELEONORA VOLLAN
EDITOR
Copyright © 2019 by Nova Science Publishers, Inc.
All rights reserved. No part of this book may be reproduced, stored in a retrieval system or
transmitted in any form or by any means: electronic, electrostatic, magnetic, tape, mechanical
photocopying, recording or otherwise without the written permission of the Publisher.
We have partnered with Copyright Clearance Center to make it easy for you to obtain permissions
to reuse content from this publication. Simply navigate to this publication’s page on Nova’s
website and locate the “Get Permission” button below the title description. This button is linked
directly to the title’s permission page on copyright.com. Alternatively, you can visit
copyright.com and search by title, ISBN, or ISSN.
For further questions about using the service on copyright.com, please contact:
Copyright Clearance Center
Phone: +1-(978) 750-8400 Fax: +1-(978) 750-4470 E-mail: [email protected].
Independent verification should be sought for any data, advice or recommendations contained in
this book. In addition, no responsibility is assumed by the publisher for any injury and/or damage
to persons or property arising from any methods, products, instructions, ideas or otherwise
contained in this publication.
This publication is designed to provide accurate and authoritative information with regard to the
subject matter covered herein. It is sold with the clear understanding that the Publisher is not
engaged in rendering legal or any other professional services. If legal or any other expert
assistance is required, the services of a competent person should be sought. FROM A
DECLARATION OF PARTICIPANTS JOINTLY ADOPTED BY A COMMITTEE OF THE
AMERICAN BAR ASSOCIATION AND A COMMITTEE OF PUBLISHERS.
Additional color graphics may be available in the e-book version of this book.
Preface vii
Chapter 1 Flocculant Polysaccharides Mainly from Plants 1
Priscilla B. S. Albuquerque, Weslley F. Oliveira,
Priscila M. S. Silva, Maria T. S. Correia
and Luana C. B. B. Coelho
Chapter 2 Coagulation and Flocculation with Plant Extracts 47
Jesús Manuel Epalza Contreras and Johan Jaramillo Peralta
Chapter 3 The Process of Water Treatment with
Aluminum Sulphate Associated with the
Application of the Cactus Opuntia Cochenillifera 69
Higgor Henrique Dias Goes, Rita de Cássia Pereira de Souza
and Joseane Débora Peruço Theodoro
Chapter 4 Flocculation: Mechanisms and Applications
for Wastewater Treatment 81
Elvis Carissimi, Cristiane Oliveira Rodrigues,
Dounia Elkhatib and Vinka Oyanedel-Craver
Chapter 5 Flocculation of AOM in Water Treatment 107
Martin Pivokonský, Jana Načeradská, Kateřina Novotná,
Lenka Čermáková and Petra Vašatová
Chapter 6 Comparison of Natural Coagulant and Chemistry in Tanning
Wastewater Treatment Using the Flocculation Process 143
Edilaine Regina Pereira, Gustavo da Silva Souza
and Joseane Débora Peruço Theodoro
vi Contents
are largely dependent on the following features: floc characteristics, flocculation kinetics,
and engineering aspects of flocculation. Finally, examples of engineering applications
using flocculation are described.
Chapter 5 - Global proliferation of algal blooms and subsequent deterioration of
water quality by organic compounds that are being produced (algal organic matter –
AOM) pose new challenges to water treatment technologies. Flocculation/coagulation
using primarily Al- and Fe-based coagulants is widely employed as an essential process
in removal of various impurities at drinking water treatment plants and is also
irreplaceable in the case of AOM elimination. This review chapter discusses current
knowledge on AOM flocculation, the impact of AOM on the removal of other
compounds and links AOM composition and character to the efficiency of flocculation,
the reaction conditions and mechanisms and finally, to the properties of flocs. In general,
the removal efficiencies of dissolved AOM are lower compared to intact phytoplankton
cells and usually reach maximum under slightly acidic pH values. The strong pH-
dependence of flocculation is attributed to the fact that the involved mechanisms are to a
great extent determined by the charge ratios in the coagulating system. Furthermore,
substantial differences in flocculation behaviour were observed between diverse AOM
constituents, i.e., between peptides/proteins versus non-proteinaceous matter and high
versus low molecular weight organics. The latter (specifically AOM < 10 kDa) are
reluctant to flocculate and would therefore require other treatment techniques. AOM has
also been reported to influence flocculation of other common impurities, both of organic
and inorganic nature. Mutual interactions have been proven, while their influence on
flocculation efficiency can be either positive or negative, depending on the AOM
character, pH conditions and on the ratio between AOM, the other polluting agents and
coagulants. Finally, AOM also appeared to alter the properties of flocs, with an impact on
the subsequent separation steps. In further research, a particular emphasis should be put
on AOM components that are difficult to coagulate, the interactions of AOM with other
impurities and on elucidation of the relationship between AOM and floc properties.
Chapter 6 - The leather tanning industry uses a large amount of toxic substances and
water. Thus, it generates effluents with high polluting load. Due to the large amount of
effluent generated by tanneries and the difficulty in their treatment, several studies have
been proposed to minimize the environmental impacts caused by inappropriate disposal
of this effluent. This study investigated the performance of natural coagulants Tanin
compared to chemical coagulants aluminium sulphate and ferric chloride commonly used
in the treatment of raw wastewater from tannery, by means of the physicochemical
processes of coagulation, flocculation and sedimentation. Using jar tests methods,
different concentrations for the coagulants (chemical and natural) were applied to the
concerned effluent and it was assessed their efficiency in removing certain parameters
such as apparent color COD, and turbity besides the behavior of electrical conductivity
and pH. The results showed that for pH and electrical conductivity parameters, there was
Preface xi
of wastewater with a high content of solids and fats and oils was evaluated. The ones that
presented a better performance for this type of wastewater are mesquite gum and
chitosan. The BPE that showed a better performance in the coagulation-flocculation test
of the wastewater with high content of fats and oils at pH 5.2, was the chitosan, with a
dose of 68 mg/L of chitosan the treated water has the following values of BOD5 = 30 mg
O2/L, fats and oil = 1 mg/L, turbidity = 2 FAU, TSS = 2 mg/L and COD = 50 mg/L.
Chapter 9 - The brewing industry presents a great generation of effluent that can
cause serious environmental impacts when not treated in a regular way since it has a high
load of organic matter in its composition. Thus, in view of the growing emergence of
breweries in Brazil and consequent increase in effluent production, alternatives are
sought for the auxiliary treatment of this effluent using coagulants. In this way, the paper
aimed to analyze the efficiency the usage of two organic and one inorganic coagulant in
the treatment of effluent from the brewing industry. The assay was performed in Jar Test,
thus simulating the coagulation, flocculation and sedimentation processes, and in
sequence the filtration process was carried out. During this process the parameters of
apparent color, electrical conductivity, pH and turbidity were monitored. Three different
treatments were used: Aluminum Sulfate (T1), Tannin (T2) and Moringa oleifera (T3).
The results showed values of apparent color removal and turbidity of more than 90% for
all treatments, but the treatment with Tanino presented the highest values of removal,
being 98.8% for turbidity removal and 99.6% for removal of apparent color. The pH
parameter showed a rise throughout the test for all treatments presenting final values
close to the neutral value, demonstrating compliance with CONAMA Resolution No.
430/2011. The electrical conductivity showed a decrease in all treatments especially in
the treatment where Tanino was used being 0.32μS.m-1 the lowest value obtained. The
use of the coagulants in question showed positive responses in the analyzed parameters in
the treatment of the effluent of the brewing industry, showing more emphasis for the
Tanino that obtained the best values. It’s an option sustainable and is possible to mix this
process with the comum process that there are in brewing industry.
Chapter 10 - Urban development contributes to increasing water pollution, so we
must develop alternative ways to treat water. Therefore, the objective of this paper was to
perform the water treatment - through the electrocoagulation process - and to calculate
the cost of the operation. The paper was done according to predefined experimental
planning that has nine samples with duplicate plus a pair of samples in central point. The
assays were carried out using a bench-level acrylic reactor (2 liters), a magnetic stirrer, a
voltage source and a monopolar electrode system arranged by means of 4 equidistantly
connected aluminum plates in parallel. The studied parameters were final pH and electric
conductivity, tested in two different voltages (5V and 11V). Sampling time were 40, 60
and 80 minutes. At the end of the trials, treatment costs were calculated to ensure that it
was applicable. The analysis of the results concluded that the electrocoagulation process
assists in the neutrality of the solution, brings the acid and basic samples near to the
Preface xiii
neutral pH (7) and the electric conductivity does not have significant change. Through the
work it was possible to observe that the eletrocoagulation process accomplished with the
continuous voltage of 5V was the one that presented the lowest cost, considering that the
energy consumption and the use of the electrode were smaller 0.03325kWh and
1.23949µg respectively.
Chapter 11 - Eutrophication, or the excessive richness of nutrients in a lake or other
body of water, is one of the most prevalent water quality problems in the United States as
well as other parts of the world. It has led to excessive growth of algal blooms, which not
only cause the death of aquatic plants and animals, but also produce high levels of toxins
and odorous compounds, thereby posing serious risks to safe drinking water and human
health. Microcystis aeruginosa is well known to be one of the most dangerous and most
common bloom-forming cyanobacteria that produces hepatotoxic microcystins. This
study evaluated the performance of the coagulation/flocculation process using aluminum
and ferric salt coagulants for the removal of both microcystins (intracellular and
extracellular) and turbidity. The effects of coagulant dosage, pH, and coagulant-aid
dosage on microcystins and turbidity removal were also evaluated. All results
demonstrated that the optimum dosage of aluminum sulphate, ferric chloride, and ferrous
sulphate for the removal of both microcystins and turbidity was found to be 60, 60, and
20 mg/L, respectively. With optimized coagulant dosages, the optimum pH value for both
microcystins and turbidity removal occurred with a pH of 7, 6, and 7 for aluminum
sulphate, ferric chloride, and ferrous sulphate, respectively. On the other hand, for all
coagulants, the optimum polyelectrolyte dosage for the removal of both microcystins and
turbidity was found to be 0.2 mg/L. Among these coagulants, aluminum sulphate (65.2%)
was found to be more effective for both microcystins and turbidity removal compared to
ferric chloride (56.3%) and ferrous sulphate (30.9%). Last, the concentration of turbidity
decreased to below 5 NTU, which is the maximum turbidity level in drinking water
determined by the World Health Organization (WHO). After coagulation/flocculation,
the total microcystin concentration (4.09 µg/L) in the water remained above the WHO
guideline value of 1 µg/L for drinking water, but the release of toxins in the water was
not observed. Therefore, both intracellular microcystin and turbidity could be removed
through the coagulation/flocculation process, whereas extracellular microcystin is still of
particular concern in drinking water treatment due to its difficult removal.
Chapter 12 - In Morocco, like all the countries of the world, socio-economic
activities coupled with population growth and changes in consumption patterns generate
a significant production of municipal solid waste. This is accompanied by a significant
increase in leachates produced during the fermentation of waste. These leachates have a
considerable impact on the environment. Moroccan regulations require leachate treatment
to reduce environmental impacts. Several techniques are currently used for the
decontamination of leachate such as (coagulation flocculation, biological treatment,
filtration, oxidation, ...). Indeed, the majority of the physicochemical treatments used for
xiv Eleonora Vollan
The effect of the order of introduction of the reagents was evaluated in terms of
elimination of turbidity, COD, color, Abs at 254 nm, phenol, detergents, ammonium ions,
nitrite, nitrate and total phosphorus.
Chapter 13 - Natural and assisted particles sedimentation are crucial operations
during the treatment of wastewater generated in many industrial processes, above all in
the wastewater treatment field. This operation is usually used as an effluent pre-
treatment; normally essential for reducing suspended solids, organic load, turbidity and
colour so as to achieve the correct clarification of wastewater. This chapter introduce the
coagulation and flocculation concepts as well as the influence of operational conditions
on assisted sedimentation. At industrial level, the factual use of both processes has been
proved on different wastewater. In this sense, this chapter outline researches about
combining or comparing assisted sedimentation (coagulation, electrocoagulation,
Preface xv
neutralization, etc.) with other operations such as oxidation processes in order to evaluate
the solids removal of the complete designed wastewater treatment focusing on OMW
treatment. In addition, in this work new trends about multifunctional and novel eco-
friendly bio-degradable flocculants and coagulants were reviewed and suggested for
OMW pretreatment.
Chapter 14 - This chapter focuses on the preparation and characterization of the
chitosan based flocculant for removal of heavy metal ion prepared from chitosan by N-
acylation with ethylenediaminetetraacetic acid monoanhydride (EDTAM). In this study, a
series of chitosan flocculants with various degree of substitution were prepared by
changing the molar ratio of EDTAM to chitosan in the preparation. The structural
characterization of the chitosan flocculants by FTIR and 13C NMR spectroscopy were
performed. The newly introduced functional group provided properties such as being a
strong chelating reagent and an amphoteric polyelectrolyte. It was found that chitosan
flocculants had good water solubility in both acidic and basic regions and precipitated in
a narrow pH region, which is close to its isoelectric point. The chitosan flocculants has an
ability to remove heavy metal ion from aqueous solution by controlling pH or initial
concentration of the flocculant, and removed Cu(II) almost completely. The flocculation
mechanism was investigated using Cu(II) and found that flocculation occurred by charge
neutralization was sensitive both pH conditions and amount of chelated Cu(II) on
chitosan flocculants. The complexation stoichiometry of Cu(II) and the EDTA residues of
chitosan flocculant determined by UV-vis titration. As a result, the chitosan flocculants
can easily remove Cu(II) from aqueous solution by only flocculation/precipitation
process, which indicated that the chitosan flocculants could be applicable for remediation
and treatment system of wastewater.
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 1
FLOCCULANT POLYSACCHARIDES
MAINLY FROM PLANTS
ABSTRACT
*
Corresponding Author Email: [email protected].
2 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
1. INTRODUCTION
because of the increasing environmental damage of this century. In addition, the scientific
community try to develop technological alternatives for this issue. The last two decades
have been marked by a crescent interest in public and scientific communities about the
use and development of biopolymers, which can be obtained from renewable sources,
displaying the important characteristic of biodegradability, with the desired
physicochemical properties of conventional synthetic materials (Albuquerque and
Malafaia, 2017; Rhim, Park and Ha, 2013).
The majority of biopolymers have rather low activity under environmental
conditions. The primary task for scientific researches, in what concerns the
polysaccharides, is to create a methodology for the synthesis of functional materials
derived from them, which have significant properties for the enhancement of their
utilization on a practical scale. This is expected to open the possibility of a real
knowledge of the nature of intermolecular interactions in aqueous solutions of
biopolymers (Oladoja et al., 2017).
The optimal choice of a particular technology for engineering of polysaccharides can
be either by chemical, physical or biological modifications. Physical (ultrasonic
disruption and microwave exposure) and biological (enzymatic degradation) changes
allow modifications of the polysaccharide molecular mass, while the chemical-one could
change the substituent types of groups, number and position (Ghimici and Nichifor,
2018).
Polysaccharide bio-based flocculants contain natural polysaccharides that are
suspected to exhibit excellent selectivity towards aromatic compounds and metals in the
absorption of particles, for example pollutants found in wastewaters (Crini, 2005).
However, it is well known that their feasibility is restricted by their physicochemical
properties, moderate flocculating activity and short shelf life. In recent years, grafted
bioflocculants have been developed and claimed due to their remarkable flocculating
ability and biodegradability. They are covalently modified by inclusion of synthetic
monomers onto their backbone to synthesise the high molecular weight grafted
copolymers that exhibit improved flocculating properties (Crini, 2005; Lee et al., 2014;
Lee et al., 2012).
The main methods for polysaccharide modification in aqueous solution can be
categorized in: (1) conventional method, (2) microwave initiated grafting method, and (3)
microwave assisted grafting method. In the conventional method, a chemical-free radical
initiator is added under an inert atmosphere to produce free radical sites on the polymer
in order to allow the addition of monomer to form the graft chain. This method is not
suitable for an industrial scale of polysaccharide production due to its low
reproducibility. The use of microwave irradiation emerges as a promising alternative; in
the microwave assisted grafting method, ions are produced by the addition of external
redox initiators to the reduction mixture, thus, the free radical initiators facilitates the
grafting reaction. In microwave initiated grafting reactions, in turn, no initiators are
Flocculant Polysaccharides Mainly from Plants 5
added and a small amount of hydroquinone is required to inhibit the grafting reactions. It
is important to mention that factors including pH, reaction temperature, reaction time,
and reactants/initiator/crosslinker dosages and ratios influence the flocculation
performance of the modified polysaccharide bio-based flocculants (Kumar, Setia and
Mahadevan, 2012; Salehizadeh, Yan and Farnood, 2017; Sen et al., 2009).
The most commonly used flocculants in industry today are inorganic and synthetic
organic flocculants; their effective flocculation activity and low cost are still preferable
over the bioflocculants, however, the question of their toxicity to human health and
environmental pollution has been a major concern (Okaiyeto et al., 2016).
Inorganic flocculants include alum, polyaluminium chloride (PAC), aluminium
chloride, aluminium sulfate, ferric chloride, ferrous sulfates or composites obtained by
the mixture of these salts. The flocculation process occurs between the salt of these
metals and the negatively charged suspended particles in a solution. This interaction leads
to a reduction in surface charge and the formation of microflocs, which in turn aggregates
to form larger flocs that can easily settle out of solution (Figure 1). Besides, its
application has caused problems of increased metal concentration or residual aluminium
in treated water, which may presents human health implications and produces large
quantity of sludge whose disposal is another problem (Lee et al., 2014).
and present polyelectrolytes in their molecular chain, which enhance their flocculating
effectiveness (Lee, Robinson and Chong, 2014; Okaiyeto et al., 2016; Suopajärvi et al.,
2013). However, problems associated to Alzheimer´s disease (Campbell, 2002) and
carcinogenic and neurotoxic risks (Rudén, 2004) to humans were already reported for
aluminium and acrylamide salt flocculants, respectively. In addition, these type of
flocculants have other limitations related to their relatively high dosage requirement, high
pH sensitivity, and poor efficiency for the coagulation of very fine particles (Sharma,
Dhuldhoya and Merchant, 2006).
alginate, gums and mucilage have been basing flocculants and attracting wide interest of
researchers (Lee, Robinson and Chong, 2014).
It is predicted that some of the active ingredients in the mucilage are responsible for
the flocculating property. Therefore, extraction becomes the essential step to isolate the
active components that exhibit the flocculating activity from the plants. There are two
methods for the production of plant-derived bioflocculants: (1) solvent extraction and
precipitation and (2) drying and grinding. Briefly, the solvent extraction and precipitation
method starts with the cleaning of the plant materials and then extraction with distilled
water overnight, followed by filtration of the mucilaginous extract and precipitation using
alcohol. In turn, the drying and grinding method starts with the cleaned materials being
dried at high temperature and then grounded and sieved to obtain the bioflocculants.
Bioflocculants prepared under solvent extraction and precipitation conditions displayed
excellent flocculating ability in the treatment of wastewater with direct flocculation
process where no coagulant and pH adjustment are required. Thus, these results suggest
that the extraction step is closely related with flocculating efficiency and plays the major
role to extract the active constituents with high flocculating activity from the plant
materials. In addition, the method used to evaluate the flocculating efficiency of the
plant-derived bioflocculants and optimise the flocculation process is called Jar Test (Lee,
Robinson and Chong, 2014; Salehizadeh, Yan and Farnood, 2017).
The most common flocculation mechanisms related for polysaccharide plant-derived
bioflocculants are charge neutralization (including electrostatic patch effects) and
polymer bridging. These mechanisms are intrinsically dependent on the adsorption of
flocculants on particle surfaces; if there is some affinity between polymer segments and a
particle surface, the adsorption process occurs (Bolto and Gregory, 2007; Lee, Robinson
and Chong, 2014; Salehizadeh, Yan and Farnood, 2017). As already mentioned in section
3, the flocculation mechanism of charge neutralisation is only applicable when the colloid
suspended particles and the added flocculants are of opposite charge. In this case, the
particle surface charge density is reduced by adsorption of the flocculant polysaccharide;
the electrostatic repulsions are reduced to a minimum. This mechanism has been found to
be quite effective for low molecular weight polysaccharide flocculants that tend to adsorb
and neutralize the opposite charges on the particles (Yang et al., 2016).
Considering the great number of neutral polysaccharides acting as bioflocculants, it is
suggested that charge neutralisation is not the mechanism responsible for flocculation. In
fact, bridging is the major mechanism in like-charged or neutral polysaccharide
flocculants, especially when they extends from the particle's surface into the solution for
a distance greater than the distance over which the interparticle repulsion occurs. In this
case, segments of the polysaccharide flocculant are adsorbed onto the particle surface
resulting in loops and tails extending into solution with the possibility of attachment of
dangling polysaccharide segments onto other adjacent particles to form flocs. Bridging
could be due to van der Waals force, static, hydrogen bonds or even chemical reaction
8 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
between some radical groups of the polysaccharide molecule and the particle; in addition,
this mechanism is known to be especially effective for high molecular weight
polysaccharide flocculants (Salehizadeh, Yan and Farnood, 2017).
Electrostatic patch and sweeping are mechanisms that contribute to the process of
flocculation. The first one is caused by polymer flocculants of high charge density
interacting with oppositely charged colloidal particles of low charge density. The net
residual charge of the polymer patch on one colloidal particle surface can adsorb onto the
oppositely charged colloidal particle. The second one forms a bulky precipitate that
enmeshed the colloidal particles, which are then either settled out or flocculate together
with the precipitate (Salehizadeh, Yan and Farnood, 2017).
Without doubts, flocculation is one of the most important and widely used treatment
process of industrial wastewaters. Many authors already highlighted the simplicity and
effectiveness of flocculation-coagulation processes (Lee, Robinson and Chong, 2014;
Okaiyeto et al., 2016; Salehizadeh, Yan and Farnood, 2017; Teh et al., 2016). In what
concerns the use of polysaccharides as bioflocculants, some characteristics could limit
their application for certain purposes. For example, chitosan is soluble in acidic media
and most of polysaccharides are water-soluble in their native form, therefore they cannot
be used as insoluble sorbents (Crini, 2005). In the flocculation process, the
polysaccharide biodegradability can lead to the loss of flocs stability and strength (Singh
et al., 2000). Also, some polysaccharides have reduced/or no bioactivity in certain
conditions (Salehizadeh, Yan and Farnood, 2017). During the time, much attention has
been paid to overcome these disadvantages by modifications of polysaccharides.
10 (Aljuboori et al., 2013; Huang et al., 2015; Li et al., 2013; Li et al., 2014; Liu et al.,
2014b).
The metabolism of microorganisms has direct relationship with cultivating
temperature. Maximum enzymatic activation can be obtained at optimal temperatures
(Zhang et al., 2007). In this way, temperature is an important parameter for bioflocculant
production and cell growth, especially considering bacteria and fungi sources.
Thermotolerant microorganisms have an optimum temperature for growth below 45 ºC,
but ability to grow at higher temperatures; they are also defined as microorganisms able
to grow at 60 ºC and below 30 ºC (Chaisorn et al., 2016).
The flocculant dosage of the polysaccharide is a main aspect for the determination of
the flocculating activity of a colloid system. When the polysaccharide dosage is lower
than the optimum, the degree of flocculation is insufficient in a colloid system that results
in stabilization and charge reversal of the colloidal particles. Thus, there is no
bioflocculant enough to adsorb onto the colloidal particle surfaces to bridge between
these particles. An overloaded dosage of the polysaccharide flocculants increases the
electrostatic repulsion forces between the colloidal particles and increases the distance
between the particles to inhibit floc formation and precipitation (Salehizadeh, Yan and
Farnood, 2017).
Figure 2. Percentage of publication in the last 10 years about the scientific literature dealing with
polysaccharides and flocculants.
Polysaccharide flocculants can be classified into three main groups based on their
source: marine, microbial, and plant flocculant polysaccharides. Their natural origin
could be related to crustacean shell wastes, seaweeds, agricultural/forestry feedstocks,
10 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
and microorganisms including bacteria, yeast, fungi, and algae. Figure 2 depicts the
percentage of published articles reporting polysaccharides and flocculants in the last 10
years.
Considering the main flocculants derived from plant-polysaccharides, we
summarized below their properties and the current researches about them.
4.1. Cellulose
content and higher tensile strength (Klemm et al., 2005). The hydroxyl groups are the
most targeted reactive groups on the cellulose main chain, which can fully or partially
react with chemical agents to obtain various derivatives with different substitution
degrees. The obtained derivatives of cellulose are been applied in food, cosmetic,
biomedical, and pharmaceutical industries, however, the application of cellulosic material
is limited due to the difficulty in processing. The high crystallinity degree and rigid
intra/intermolecular hydrogen bonds result in its insolubility in water and most organic
solvents. In addition, cellulose becomes amorphous in water at 320 °C and 25 MPa and
can be converted chemically into its monomeric units by reacting with concentrated acids
at high temperatures (Deguchi, Tsujii and Horikoshi, 2006; Liu, Willför and Xu, 2014;
Zhang, Lin and Yao, 2015).
Problems associated to processing cellulose and its limited solubility are observed for
other natural polymers since they present increasingly application in the industrial
technology. Most polymers do not present biological activity until modifications are
made, so it is important to mention that numerous attempts are being performed to
minimize these certain drawbacks with chemical modifications in the polymeric structure
(Albuquerque et al., 2017). Cellulose can be considered an efficient alternative to
produce environmentally friendly functional materials and chemicals such as
bioflocculants due to its inherent physicochemical characteristics, which are improved by
chemical modifications (Salehizadeh, Yan and Farnood, 2017).
The most recent researches on bioflocculant production are reporting isolated strains
(Lee and Chang, 2018; Shahadat et al., 2017). The unique potential of bioflocculants
produced from microorganisms was first investigated in Levure casseeuse yeast by Louis
Pasteur, and a similar trend in bacterial culture was also observed by Bordet in 1899
(Shahadat et al., 2017). More recently, it was reported that some bacteria can produce
bioflocculants by utilizing biomass from raw materials: a lignocellulose-degrading strain
Cellulosimicrobium cellulans L804 isolated from corn farmland soil presented the ability
to produce bioflocculants by the degradation of lignocellulosic biomass (Liu et al., 2015).
Liu et al. (2017a) reported an alkaliphilic strain Bacillus agaradhaerens C9 bioflocculant
producer by using untreated rice bran as carbon source, and Guo et al. (2018) reported a
novel cellulase-free xylanase-producing bacterium G22 with the potential ability to
directly convert various biomasses to bioflocculants.
The scientific literature dealing with polysaccharide flocculants derived from plants
is still scarce when compared to the works reporting bioflocculants produced from
microorganisms, or even those one reporting commercial polysaccharide bioflocculants.
However, the importance of these kind of studies rises with the increasingly serious
environmental problems especially associated to the discharge of effluents. For example,
Liu et al. (2014a) developed an efficient and eco-friendly flocculant from Phyllostachys
heterocycle bamboo pulp cellulose grafted with polyacrylamide for an effluent from
paper mill. After that, Zhu et al. (2016) optimized the method and employed the co-
12 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
polymer with the metal ions Fe3+, Al3+ or Ca2+ to treat the effluent from a surfactant
manufacturer. Liimatainen et al. (2012) evaluated the effectiveness of a flocculation
treatment based on alum and soluble or nanoparticular anionic derivatives of dialdehyde
cellulose derived from the bleached birch chemical wood pulp obtained in dry sheets of
Betula verrucosa and B. pendula, by studying the removal of colloidal material in a
model suspension containing kaolin.
4.2. Starch
methods and high cost of the cationic monomers. For example, three different starch-
based flocculants, all of them containing the strongly cationic quaternary ammonium salt
groups but at different positions, have been designed and prepared through etherification,
graft copolymerization, or their combination. The effects of chain architectures and
charge properties on flocculation of humic acid (HA) and its floc properties have been
investigated in terms of several environmental parameters including pH, dose, and initial
HA concentration (Wu et al., 2016). Posteriorly, authors obtained four variations with
different cationic contents of two versions of the starch-based flocculants reported in the
previous work. The environmental parameters pH and flocculant dose were evaluated
with the effects of structural factors, i.e., cationic group contents and distributions, on the
flocculation of the hairwork effluents (Du et al., 2017). These starch-based flocculants
were also evaluated by the efficient flocculation of real secondary textile dyeing
effluents. Starch-graft-poly[(2-methacryloyloxyethyl) trimethyl ammonium chloride]
(STC-g-PDMC) and starch-3-chloro-2-hydroxypropyl trimethyl ammonium chloride
(STC-CTA) have been systematically investigated and compared with that of the
traditional inorganic coagulant polyaluminum chloride, and the obtained results are of
significance in guiding the design and selection of a suitable polymeric flocculant in
treating target wastewater (Wu et al., 2017a).
Liu et al. (2017b) prepared a series of cationized starch-based flocculants (ST-CTA)
containing various quaternary ammonium salt groups on the starch backbone by using a
simple etherification reaction. They observed an effective performance for the
flocculation of kaolin suspension, two bacterial suspensions (of Escherichia coli and
Staphylococcus aureus), and two contaminant mixtures (kaolin and each bacterium)
under most pH conditions. Wang et al. (2013) prepared a cationic grafted starch (ST-g-
PDMC) with high flocculation performance for kaolin suspensions and efficient
dewatering of anaerobic sludge. More recently, Huang et al. (2017) reported that the St-g-
PDMC was efficient for the flocculation of kaolin suspensions and for the inhibition of E.
coli by introducing new quaternary ammonium groups. A grafted amphoteric starch-
based flocculant (carboxymethyl-starch-graft-aminomethylated-polyacrylamide, CMS-g-
APAM) was efficiently developed by Huang et al. (2016), in which the cationic groups
were randomly distributed on the polyacrylamide branched chains using Mannich
reaction. A series of cationic starches were developed by incorporating a cationic moiety
(STC-CTA) onto the backbone of starch in presence of NaOH. The flocculation
characteristics of these starches were evaluated in silica suspension and compared with
various commercially available flocculants by jar test (Pal, Mal and Singh, 2005).
It is important to mention that the majority of researches about starch flocculants deal
with the commercial presentation of the polysaccharide. Well-prepared natural
flocculants, with or without chemical modifications, owing to their superior performance
and environmental friendliness, have wide-ranging uses in wastewater treatment,
especially when more effective production techniques are developed and optimized in the
14 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
near future. Additionally, it is important to highlight the volume of the sludge produced
from the treatment associated to them and the considerable reduction in the cost of sludge
disposal.
4.3. Pectin
Pectins are ubiquitous plant polysaccharides. They are present in the cell walls
located in the middle lamella, and primary and secondary cell wall. The chemical
structure of this water-soluble gum is heterogeneous, depending on the origin, location in
the plant and extraction method, being composed at least of 65% of galacturonic acid
units plus rhamnose, arabinan, galactan and arabinogalactan (Müller-Maatsch et al.,
2016; Tamnak et al., 2016). Actually, the uronic acid residues linked through a-1-4-
glycosidic bonds make the main structural polysaccharide motif, while there exist three
pectin’s domains: α-(1-4)-linked linear homogalacturonic backbone (HG) alternating
with two types of highly branched rhamnogalacturonans regions called RG-I and RG-II.
The first region (RG-I) is substituted with side chains of arabinose and galactose units,
while the second one (RG-II) has a highly conserved structure, consisting of the HG
backbone branched with eleven different monosaccharides, including some rare sugars
such as apiose, aceric acid, 2-O-methylxylose, 2-O-methylfucose, 2-keto-3-deoxy-d-
manno-octulosonic acid, and 3-deoxy-d-lyxo-2-heptulosaric acid. In all natural pectins,
some of the carboxyl groups exist in the methyl ester form (Albuquerque et al., 2017;
Liu, Willför and Xu, 2015).
Pectin can be categorized as anionic polysaccharides mainly derived from the cell
wall of plants, and also from food-industrial wastes of fruits, whose main sources are
citrus peels (e.g., lemon, lime, orange, and grapefruit), apple pomace, and sugar beet pulp
(Babbar et al., 2015; Marić et al., 2018). Moreover, pectin from non-conventional sources
has been evaluated, for example, from cocoa husks (Chan and Choo, 2013), mulberry
branch bark (Liu, Jiang and Yao, 2011), jackfruit peel (Xu et al., 2018), faba bean hulls
(Korish, 2015), sisal waste (Yang et al., 2018), watermelon rind (Maran et al., 2014),
pomegranate peels (Pereira et al., 2016), potato pulp (Yang, Mu and Ma, 2018), Ubá
mango peel (Oliveira et al., 2018), pistachio green hull (Chaharbaghi, Khodaiyan and
Hosseini, 2017), pequi peel (Leão et al., 2018), and banana peel (OLiveira et al., 2016).
In the industry, the major pectin component (homogalacturonan) can be obtained using
hot water or chemically by acid extraction, but some innovative extraction approaches
have been developed to improve extraction process and pectin quality (Salehizadeh, Yan
and Farnood, 2017).
Pectin structure determines in a great manner its physicochemical properties and
applications; for example, pectin is efficiently used by the food industry due to their
ability to form gels under certain circumstances and to increase the viscosity of drinks. It
Flocculant Polysaccharides Mainly from Plants 15
is also widely applied as stabilizers in acid milk products, and some may have other
pharmaceutical uses (Müller-Maatsch et al., 2016).
Pectin has many applications because of its low-cost, read availability, harmless and
green nature. However, it may not provide the proper emulsifying activity because of its
hydrophilicity. The following modification techniques have been recommended for the
improvement of pectin physicochemical and functional properties: chemical modification
(such as saponification, distillation, and esterification), enzymatic modification (by pectin
methylesterases), and physical modification (by heat treatment and microwave) (Tamnak
et al., 2016).
Yokoi et al. (2002) were one of the first authors reporting pectin as an efficient
flocculant for various suspensions. They found that pectin had a flocculating activity,
however, in that time; there has so far been no report about flocculating activity of pectin
and its applicability as a flocculating agent. Their results demonstrated that pectin had
flocculating activity in a kaolin suspension, and this activity was enhanced by the
addition of Al3+ and Fe3+ to the suspension; in relation to organic suspensions such as
cellulose and yeast, pectin acted as flocculant when 0.1–0.2 mM Fe3+ was present in the
suspensions. Other inorganic suspensions of activated carbon and acid clay were
flocculated by pectin in the presence of Al3+ or Fe3+.
A commercial citrus pectin with 60% esterification and the common organic
synthetic flocculant polyacrylamide were characterized and used to optimize the
treatment processes of both flocculants in synthetic turbid waste water. The results
reported by Ho et al. (2010) were important for the industrial application of pectins since
the main concern for industry is to use low flocculant concentrations to achieve
maximum results. In this case, the usage of pectin achieved that goal and proved to be
effective at a low concentration of 3 mg/L. The influence of four commercial citrus
pectins, with different degrees of esterification, and pectin extracted from pomelo (Citrus
maxima) on the stability of indomethacin suspension was extensively investigated by
Piriyaprasarth and Sriamornsak (2011). The results demonstrated that the extracted pectin
had comparable activity to the commercial ones. Moreover, the use of low concentration
of pectin and ferric ions allowed obtaining indomethacin suspensions with suitable
stability and redispersibility. More recently, pectin extracted from Nopal (Opuntia ficus-
indica) was tested to treat synthetic waste water contaminated with metallic ions (Ibarra-
Rodríguez et al., 2017). Authors well-characterized the pectin and then demonstrated its
effectiveness in the removal of the heavy metals by coagulation-flocculation treatment.
It is possible to observe that flocculation properties of pectin have been confirmed by
the scientific literature. In fact, pectin can be utilized as a harmless bioflocculant, since it
is biodegradable, and edible and non-toxic toward humans and the environment;
however, the publications are still discreet when compared to the most popular
polysaccharide bio-based flocculants.
16 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
and a sludge dewatering performance more efficient than inorganic flocculants (Liu et al.,
2014b).
An exopolysaccharide produced by cianobacterium Anabaena sp. BTA992 during its
photoautotrophic growth was evaluated and it was observed bioflocculant activity with
potential properties for commercial production and industrial applications (Khangembam
et al., 2016).
A biomass-degrading bacterium Pseudomonas sp. GO2 was able to produce a low-
cost bioflocculant from untreated corn stover through directly hydrolyzing biomasses.
Biochemical analysis showed a composition of polysaccharides (59%) with uronic acid
(34.2%), protein (32.1%) and nucleic acid (6.1%), and a high flocculation potential
flocculating to harvest the green microalgae Chlorella zofingiensis and Neochloris
oleoabundans (Guo et al., 2017).
Achromobacter xylosoxidans TERI L1 isolated from oil refinery waste produces an
exopolysaccharide bioflocculant encompassed of various functional charged groups with
flocculating activity for dispersed kaolin clay particles in suspension and to adsorption of
multi metals from environment (Subudhi et al., 2016). Thus, this flocculant is a good
candidate for heavy metal removal in contaminated waste-water.
Green synthesis of silver nanoparticles was a method developed based in a
polysaccharide flocculant produced by the marine actinobacterium Streptomyces sp.
MBRC-91. The biosynthesized silver nanoparticles showed strong antibacterial potential
in sewage water, being useful for wastewater treatment (Manivasagan et al., 2015). A
flocculant exopolysaccharide from Arthrobacter sp. B4 was also used for green synthesis
of silver nanoparticles, resulting in nanoparticles highly stable, with antimicrobial activity
and low phytotoxicity (Yumei et al., 2017). This represents the use of microbial
flocculants as a potential tool in many fields, including the medical therapies.
Conventionally, fungi play valuable roles in environment and are involved in global
biotechnological processes, including the production of enzymes, polysaccharides, lipids
and pigments. Some fungal species have been reported as bioflocculant polysaccharide
producers with high flocculating activities.
An oleaginous fungal Curvularia sp. strain DFH1 demonstrated its potential to
coproduce lipids and an exopolymer with a powerful flocculant activity of 95% (Abul-
Elreesh and Abd-El-Haleem, 2014). The chemical analysis of the exopolymer shows a
composition of protein and polysaccharide. An analysis of infrared spectrum indicates the
presence of carboxyl and hydroxyl groups, typical in sugar derivatives.
A polysaccharide bioflocculant named IH-7 was produced and isolated from
Aspergillus flavus fermentation medium (Aljuboori et al., 2013). The bioflocculant
22 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
consisted basically of protein (28,5%) and sugar (69,7%), including neutral sugar, uronic
acid and amino sugar, and showed a good flocculating activity in kaolin suspension
without cation addition.
Two bioflocculant exopolysaccharides were isolated from fungal cultures named
SGMP1 and SGMP2 found in soil samples collected from New Vallabh Vidyanagar,
Gujarat (India). The flocculating activity was 99%, and the fungal isolates could remove
the cations Al3+ and Fe3+ in ideal concentrations for application in bioremediation of
heavy metals (Patel et al., 2014).
A bioflocculant (MBFA18) produced by Aspergillus niger (A18) isolated from soil
sample using potato starch wastewater as nutrients showed strong flocculant efficiency,
less dosage, sludge amount and moderate treating condition comparable to the chemical
flocculants. A flocculating rate of approximately 90% was achieved for kaolin clay under
optimal cultivation condition (Pu et al., 2018).
Exopolysaccharides from marine algae have been explored due to their potential
applications, including as bioflocculant agent, considered more efficient that commercial
flocculants.
Alginates are polysaccharides firstly isolated from brown seaweeds (Phaeophyta),
and recently identified as produced also by two genera of soil bacteria, Pseudomonas and
Azotobacter (Hay et al., 2013). This polysaccharide is constituted to over than 200
different types varying in length and array of β-D-mannuronic acid (M) and α-L-
guluronic acid (G) monomer units (Albuquerque et al., 2017).
Alginates are biocompatible, biodegradable, and non-immunogenic biopolymer
polyelectrolytes (Yang, Xie and He, 2011). The physical properties in aqueous medium
for alginates depend not only on the M/G ratio, but also on the distribution of M and G
units along the chain. At this moment, it is important to highlight that the main property
of alginates is their ability to retain water. Because of their linear structure, and high
molecular weight, alginates form strong films and good fibres in the solid state. Their
gelling and stabilizing properties are also very important characteristics for alginates; the
stiffness of the alginate chains and the complex with counterions could be attributed to
the composition (M/G ratio) and distribution of M and G units in the chains. The higher
content of G units form stable crosslinked junctions with divalent counter ions (for
example, Ca, Ba, and Sr, unless Mg), so the crosslinked network can be considered a gel.
In addition, the low pH also forms acidic gels stabilized by hydrogen bonds
(Albuquerque et al., 2017; Rinaudo, 2008; Yang, Xie and He, 2011).
Flocculant Polysacharides Mainly from Plants 23
5.4.1. Chitosan
According to Wu et al. (2012), chitin is the second most abundant polymer after
cellulose. It is widely synthesized by many living organisms such as fungi, yeasts, algae
and squid pen, being found in exoskeletons of crabs, lobsters, shrimp and crab shells, as
well as insect cuticles. Depending on its source, chitin occurs mainly as two allomorphs,
namely α (the most abundant) and β forms. The chemical structure (poly-β-(1→4)-N-
acetyl-D-glucosamine) of chitin can be partially degraded by acid to obtain a series of
24 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
Different types of industries, such as textiles and plastic, use dyes to colouring their
products, generating large volumes of colored wastewater. However, dyes are substances
capable of absorbing ultraviolet light and consequently decrease this absorption by
photosynthetic organisms and reduce dissolved oxygen content of aquatic environment;
characterizing, thus, a serious environmental pollutant (Rangabhashiyam et al., 2013;
Bello et al., 2017). Coagulation-flocculation has been considered a cost effective
technology with excellent color removal ability (Verma et al., 2012).
A polysaccharide with cationic behaviour that is widely used for removal of
negatively charged dyes in wastewater is the chitosan; anionic dyes can be adsorbed
electrostatically with the protonated amine groups of chitosan (Lee et al., 2014; Vakili et
al., 2014). For example, about 99% of the acid blue 92 dye was removed by chitosan,
allowing recovery of the biopolymer from formed flocs with the dye using 0.1 M NaOH
and with subsequent reuse of chitosan in acetic acid solution (Szygula et al., 2009).
Additionally, modifications in chitosan structure as well as its association with other
polymers have been increasingly studied to improve flocculant action in dye removal.
A chitosan-based cationic polymer was elaborated through the modification of this
polysaccharide by (3-chloro-2-hydroxypropyl) trimethylammonium chloride; such
26 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
chitosan modification was useful for dye melanoidin removal from industrial wastewater
(Momeni et al., 2018). In an attempt to elaborate amphoteric flocculants, capable of
eliminating both anionic and cationic dyes, Yang et al. (2013) prepared grafting
flocculants, formed by carboxymethyl chitosan-graft-polyacrylamide (CMC-g-PAM).
These authors have found that the ability to remove methyl orange (anionic dye) and
basic bright yellow (cationic dye) initially occurs by charge neutralization between
produced flocculant and dyes, but flexibility of PAM graft chains allowed formation of
large and insoluble flocs with net-like structure, allowing an easy sedimentation in water
treatment (Yang et al., 2013). The synthetic monomer of acrylamide was used to graft
chitosan and lignin (CAMCL) forming a ternary graft copolymer, since these two
polymers together provide more functional groups to attract dyes, such as orange C-3R,
methyl orange and acid black-172. In addition, flocculating mechanism of CAMCL
occurred by the charge neutralization, bridging and sweeping effects (Cui et al., 2017;
Lou et al., 2018).
Cellulose has a regular structure and matrix formed by hydroxyl groups presenting a
high degree of polymerization and susceptible to undergo chemical modifications making
it a relevant flocculant (Kono and Kusumoto, 2015; Ferreira et al., 2016). Quaternized
celluloses and ampholytes of this polysaccharide in different degree of cationic
substitution, for example, had flocculating capacity against anionic dyes, such as acid red
13 and acid blue 92 (Kono and Kusumoto, 2015; Kono, 2017). However, carboxymethyl
cellulose undergoing the grafting process with hydrolyzed polyacrylamide, generating
CMC-g-HPAM, was able to remove the methylene blue cationic dye. The highest
removal efficiency occurred under alkaline conditions where the carboxyl groups of
CMC-g-HPAM were deprotonized, forming -COO- groups, allowing electrostatic
interactions with methylene blue, favoring such flocculation (Cai et al., 2013).
Gum polysaccharides also went through the grafting process, such as the ghatti-
crosslinked-polyacrylamide (Gg-cl-PAAM) hydrogel synthesized capable of adsorbing
and removing cationic (rhodamine B and brilliant green) and anionic (methyl orange and
congo red) dyes (Mittal et al., 2018). Xylan, a xylose polymer, is another polysaccharide
which, when phosphorylated, acted as a flocculant removing more than 95% of cationic
ethyl violet dye (Liu et al., 2018). A macromolecule originated using corn ethanol
wastewater, named compound biopolymer flocculant (CBF), has in its composition more
than 80% polysaccharide; CBF presented methylene blue removal, whose flocculation
process occurs primarily by adsorption bridging and charge neutralization by its polar
functional groups (Xia et al., 2018). The exopolysaccharide (EPS) produced and purified
from Paenibacillus elgii B69, was able to promote high rate of decolourization for
methylene blue and red X-GRL positive dyes, while less than 50% for anionic and
neutral dyes (Li et al., 2013).
Flocculant Polysacharides Mainly from Plants 27
Heavy metal ions, such as lead, copper and zinc, can contaminate water due to
natural processes, since such elements can be stored in the soil and transported by surface
waters. In addition, anthropogenic sources contribute to this type of pollution, such as
industrial activities including mining and coal combustion (Kobielska et al., 2018). These
ions are extremely toxic and can cause serious human health problems, such as damage to
the neurological system, blood disorders, toxicity in different organs and even death
(Shtenberg et al., 2015). In front of this serious scenario of contamination with heavy
metal ions, to remove these contaminants in wastewater is very important to leave it free
or in non-toxic concentrations of these ions; such removal can be carried out through
bioflocculants, which include polysaccharides (Salehizadeh and Yan, 2014).
Some polysaccharides, such as xanthan gum and starch, possess many –OH groups
and act as metal ion coordination sites; while other polysaccharides, for example,
chitosan with –NH2 groups and sodium alginate apart from –OH groups –COO- can bind
to the metal ions (Kolya and Tripathy, 2013). Chitosan, for example, promoted
simultaneous adsorption of heavy metal ions (iron, nickel and copper) and salt anion
(sulfate). At pH range 5 to 6, chitosan has protonated and unprotonated amine groups, in
which the latter group form chelate complexes with salt cations and salt anions
are adsorbed through electrostatic interaction on chitosan (Mende et al., 2016).
Polysaccharide inulin underwent a carboxymethylation process to synthesize
carboxymethyl inulin (MIC), which removed metal ions in a kaolin suspension, reducing
the total iron, chromium VI and manganese (II) content. The mechanism proposed for
this chelation is the interaction with the cations by oxygen of the ether linkage and by
oxygen of the carboxylate group through five membered ring formation (Rahul et al.,
2014). A pectin, extracted from O. ficus-indica had affinity to the different metallic ions
(Ca2+, Cu2+, Zn2+, Cr3+, Ni2+, Pb2+ and Cd2+); since pectin is a linear molecule composed
of galacturonic acid, whose hydroxyl and carboxylic groups can be responsible for this
affinity with the metal ions studied (Ibarra-Rodríguez et al., 2017).
Flocculating power of polysaccharides produced by different bacteria also has been
determined. EPS produced by bacteria, such as cyanobacteria, has negative charges and
can act as a chelating agent of heavy metal ions with positive charge (Bhunia et al.,
2018). The EPS bioflocculant purified from P. elgii B69 also had high capacity to adsorb
Al3+, followed by Pb2+, Cu2+ and Co2+ (Li et al.; 2013). A flocculant system was made
using EPS from Rhizobium sp. and polyethyleneimine (PEI), which was able to remove
Cu2+. It was proposed that such process occurred firstly by binding Cu2+ to negatively-
charged groups, mainly carboxylate, of the EPS followed by link of the EPS-Cu2+
complex with PEI resulting in firm and large flocs with easy removal (Escobar et al.,
2015). While more than 90% of heavy metals (Zn2+, Cd2+, Pb2+, Cu2+ and Ni2+) were
adsorbed to the EPS of glycoprotein nature produced by Achromobacter xylosoxidans
28 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
when grown in culture medium supplemented with these ions (Subudhi et al., 2016). A
polysaccharide molecule (MSI021) was purified from Bacillus cereus; and characterized
as a bioflocculant heavy metal remover. A solution containing heavy metal salts (HgCl2,
ZnCl2 and CuSO4) with a bioluminescent bacterium, Vibrio harveyi, was treated with
MSI021. The presence of MSI021 in the solution allowed the V. harveyi growth, which
continued with its luminescent phenotypic expression, differently from the absence of
MSI021, in which there was a decrease in the luminescence of this bacterium; such
phenomenon was possible due to this bioflocculant molecule that mitigated heavy metal
toxicity (Sajayan et al., 2017).
About 42% of global industrial wastewater is produced by pulp and paper industry.
The processes in the paper manufacturing involve pulping, bleaching and papermaking;
in fact, in all these steps effluents composed of different chemical species are generated
(Toczyłowska-Mamińska, 2017). Wastewater from pulp and paper industry have a
diversity of organic and inorganic contaminants, such as resin acids, fatty acids,
adsorbable organic halide (AOX) and total suspended solids (TSS) besides being able to
have high content of biological oxygen demand (BOD) and chemical oxygen demand
(COD) (Ashrafi et al., 2015; Kumar et al., 2015). These undiluted untreated effluents are
toxic to aquatic organisms and their treatment can be carried through the flocculation
physicochemical method, which includes the flocculating polysaccharides (Kumar et al.,
2015).
Thus, a copolymer flocculant prepared by grafting (2-methacryl-oyloxyethyl)
trimethyl ammonium chloride onto chitosan (chitosan-g-PDMC) has been used to treat
pulp mill wastewater. This novel cationic chitosan-based flocculant promoted efficient
removal of COD, turbidity and lignin as well as water recovery. Grafting of chitosan
showed in favour of the double electric layer compression, charge neutralization and
improved the sweep-floc effects (Wang et al., 2009). In another work, raw and undiluted
pulp and paper mill effluent was treated with Cassia obtusifolia seed gum, which
removed COD and TSS an up to 36.2 and 86.9%, respectively; and the flocs formed after
the treatment had a morphology highly fibrous-like and aggregate (Subramonian et al.,
2014). Polysaccharides from potato starch have been modified by benzylation and
insertion of hydroxypropyl-trimethyl ammonium (HPMA) moieties of different
substitution degrees, generating benzyl 2-hydroxypropyl-trimethylammonium starch
chloride (BnHPMAS). Flocculation performance was tested for BnHPMAS in a model of
paper industry wastewater; these amphiphilic starch derivatives removed contaminants
from recycled paper processing, so named stickies, including total organic carbon (TOC)
(Genest et al., 2015). A cationic bioflocculant has been produced through polymerizing
Flocculant Polysacharides Mainly from Plants 29
More than 300 million ton of organic waste are produced each year by agricultural
and food processing industries in the United States; sustainable treatment of such waste
has become a worldwide challenge (Sheets et al., 2015). These organic substances found
in wastewater from such industries comprise TSS, COD, organic colloids, sludge,
oil/grease, beyond dissolved inorganics (Teh et al., 2016). In addition, nitrogenous
compounds, especially ammonium nitrogen, can be present in large quantities in food
industry wastewater (Zhukova et al., 2011). Phosphate is a common constituent of
agricultural fertilizers and such chemical specie can contaminate effluents; excess of
phosphorus in the aquatic environment can contribute to the process of eutrophication
(Adesoye et al., 2014). Therefore, agricultural and food contaminants in wastewater need
to be removed, and such process can be performed by flocculant polysaccharides.
Wastewater is generated in potato starch process that is considered one of the most
polluted in food industry. Potato starch wastewater (PSW) has been treated with a
bioflocculat complex denominated MBF917, produced by Rhizopus sp. M9 and M17,
such macromolecule is composed of more polysaccharide than protein. MBF917
removed the turbidity and COD from the PSW (Pu et al., 2014). A bioflocculant has been
extracted from Paenibacillus polymyxa, and such compound showed to be thermostable
with indicative of its main backbone being formed by polysaccharide; this bioflocculant
was still used to treat PSW by reducing COD and turbidity efficiently (Guo et al., 2015a;
Guo et al., 2015b). Another bioflocculant of polysaccharide nature has been produced
30 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
CONCLUSION
The purpose of this review was to approach scientific literature dealing with
polysaccharide bio-based flocculants with special emphasis for those derived from plants.
Moreover, this review emphasized aspects such as extraction, purification, modification,
characterization, and the broad range of applications of natural flocculants in industry.
The most current scientific publications demonstrate the industrial concern for the
use of new substrates, besides the reduction in the costs of bioflocculants´ production and
the constant awareness about the substitution of inorganic flocculants for the innovative
ones. It was remarkable to note the preference of bioflocculants in comparison with
inorganic flocculants due to their low cost, availability, biodegradability and biosafety; in
addition, they can be obtained from renewable resources and their application is directly
related to the improvement of quality of life.
Several studies investigated the flocculating activity of polysaccharide bio-based
flocculants in wastewater treatment; they are technically promising as flocculants with
high removal efficiency of suspended solids, total dissolved solids, turbidity, colour and
dye. However, its development is limited with variation of flocculating efficiency, short
shelf life, and high production cost.
The scientific literature dealing with flocculants polysaccharides derived from plants
is still scarce when compared to bioflocculants produced from microorganisms, or even
those reporting commercial polysaccharide bioflocculants. Regarding issues such as
Flocculant Polysacharides Mainly from Plants 31
biosafety and the sake of ecology, more qualitative and quantitative research are
necessary for further exploitation and applications of polysaccharide bio-based
flocculants in different industries.
REFERENCES
Ashogbon, A. O. and Akintayo, E. T. (2014) Recent trend in the physical and chemical
modification of starches from different botanical sources: A review. Starch/Staerke,
66(1–2), 41–57.
Ashrafi, O., Yerushalmi, L. and Haghighat, F. (2015) Wastewater treatment in the pulp-
and-paper industry: a review of treatment processes and the associated greenhouse
gas emission. Journal of Environmental Management, 158, 146-157.
Babbar, N., Dejonghe, W., Gatti, M., Sforza, S. and Elst, K. (2015) Pectic
oligosaccharides from agricultural by-products: production, characterization and
health benefits. Critical Reviews in Biotechnology, 36(4), 594-606.
Bello, O. S., Lasisi, B. M., Adigun, O. J. and Ephraim, V. (2017) Scavenging rhodamine
B dye using Moringa oleifera seed pod. Chemical Speciation & Bioavailability,
29(1), 120-134.
Bhunia, B., Prasad Uday, U. S., Oinam, G., Mondal, A., Bandyopadhyay, T. K. and
Tiwari, O. N. (2018) Characterization, genetic regulation and production of
cyanobacterial exopolysaccharides and its applicability for heavy metal removal.
Carbohydrate Polymers, 179, 228-243.
Bolto, B. and Gregory, J. (2007) Organic polyelectrolytes in water treatment. Water
Research, 41(11), 2301–2324.
Brinchi, L., Cotanaa, F., Fortunatib, E. and Kenny, J. M. (2013) Production of
nanocrystalline cellulose from lignocellulosic biomass: Technology and applications.
Carbohydrate Polymers, 94(1), 154–169.
Busi, S., Karuganti, S., Rajkumari, J., Paramanandham, P., and Pattnaik, S. (2016) Sludge
settling and algal flocculating activity of extracellular polymeric substance (EPS)
derived form Bacillus cereus SK. Water and Environment Journal, 31(1), 97-104.
Cai, T., Yang, Z., Li, H., Yang, H., Li, A. and Cheng, R. (2013) Effect of hydrolysis
degree of hydrolyzed polyacrylamide grafted carboxymethyl cellulose on dye
removal efficiency. Cellulose, 20(5), 2605-2614.
Campbell, A. (2002) The potential role of aluminium in Alzheimer’s disease.
Nephrology, dialysis, transplantation : official publication of the European Dialysis
and Transplant Association - European Renal Association, 17(2), 17–20.
Chaharbaghi, E., Khodaiyan, F. and Hosseini, S. S. (2017) Optimization of pectin
extraction from pistachio green hull as a new source. Carbohydrate Polymers, 173,
107–113.
Chaisorn, W., Prasertsan, P., O-Thong, S. and Methacanon, P. (2016) Production and
characterization of biopolymer as bioflocculant from thermotolerant Bacillus subtilis
WD161 in palm oil mill effluent. International Journal of Hydrogen Energy, 41(46),
21657–21664.
Chakrabarti, S., Banerjee, S., Chaudhuri, B., Bhattacharjee, S. and Dutta, B. K. (2008)
Application of biodegradable natural polymers for flocculated sedimentation of clay
slurry. Bioresource Technology, 99(8), 3313–3317.
Flocculant Polysacharides Mainly from Plants 33
Chan, S. Y. and Choo, W. S. (2013) Effect of extraction conditions on the yield and
chemical properties of pectin from cocoa husks. Food Chemistry, 141(4), 3752–3758.
Chen, P., Cho, S. Y. and Jin, H. J. (2010) Modification and applications of bacterial
celluloses in polymer science. Macromolecular Research, 18(4), 309–320.
Chen, X., Si, C. and Fatehi, P. (2018) Cationic xylan- (2-methacryloyl-oxyethyl trimethyl
ammonium chloride) polymer as a flocculant for pulping wastewater. Carbohydrate
Polymers, 186, 358-366.
Crini, G. (2005) Recent developments in polysaccharide-based materials used as
adsorbents in wastewater treatment. Progress in Polymer Science (Oxford), 30(1),
38–70.
Cui, B., Li, L., Zeng, Q., Lin, F., Yin, L., Liao, L., Huang, M and Wang, J. (2017) A
novel lectin from Artocarpus lingnanensis induces proliferation and Th1/Th2
cytokine secretion through CD45 signaling pathway in human T lymphocytes.
Journal of Natural Medicines, 71(2), 409-421.
Cui, G., Wang, X., Xun, J. and Lou, T. (2017) Microwave assisted synthesis and
characterization of a ternary flocculant from chitosan, acrylamide and lignin.
International Biodeterioration and Biodegradation, 123, 269–275.
Czemierska, M., Szczes, A., Holysz, L., Wiater, A., Jarosz-Wilkolazka, A. (2017)
Characterisation of exopolymer R-202 isolated from Rhodococcus rhodochrous and
its flocculating properties. European Polymer Journal, 88, 21-33.
Dao, V. H., Cameron, N. R. and Saito, K. (2016) Synthesis, properties and performance
of organic polymers employed in flocculation applications. Polymer Chemistry, 7(1),
11–25.
Deguchi, S., Tsujii, K. and Horikoshi, K. (2006) Cooking cellulose in hot and
compressed water. Chemical Communications, 31, 3293
Dharani, M. and Balasubramanian, S. (2015) Synthesis and characterization of chitosan-
g-N-methyl piperazinium chloride: A hybrid flocculant. International Journal of
Biological Macromolecules, 81, 778–784.
Du, H., Yang, Z., Tian, Z., Huang, M., Yang, W., Zhang, L. and Li, A. (2018) Enhanced
removal of trace antibiotics from turbid water in the coexistence of natural organic
matters using phenylalanine-modified-chitosan flocculants: Effect of flocculants’
molecular architectures. Chemical Engineering Journal, 333, 310–319.
Du, Q., Wei, H., Li, A. and Yang, H. (2017) Evaluation of the starch-based flocculants on
flocculation of hairwork wastewater. Science of the Total Environment, 601–602,
1628–1637.
Escobar, E. C., Navarro, R. R., Nayve, F. R. P., Borines, M. G. and Ventura, J. S. (2015)
Copper (II) removal from industrial effluent using a coagulation-flocculation process
employing Rhizobium extracellular polysaccharide. Philippine e-Journal for Applied
Research and Development, 5, 1-10.
34 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
Hossain, M., and Mondal, I. H. (2014) Biodegradable surfactant from natural starch for
the reduction of environmental pollution and safety for water living organism.
International Journal of Innovative Research in Advanced Engineering, 1(8), 424–
433.
Huang, M., Liu, Z., Li, A. and Yang, H. (2017) Dual functionality of a graft starch
flocculant: Flocculation and antibacterial performance. Journal of Environmental
Management, 196, 63–71.
Huang, M., Wang, Y., Cai, J., Bai, J., Yang, H. and Li, A. (2016) Preparation of dual-
function starch-based flocculants for the simultaneous removal of turbidity and
inhibition of Escherichia coli in water, Water Research, 98, 128-137.
Huang, X., Gao, B., Yue, Q., Zhang, Y. and Sun, S. (2015) Compound bioflocculant used
as a coagulation aid in synthetic dye wastewater treatment: The effect of solution pH.
Separation and Purification Technology, 154, 108–114.
Ibarra-Rodríguez, D., Lizardi-Mendoza, J., López-Maldonado, E. A. and Oropeza-
Guzmán, M. T. (2017) Capacity of ‘nopal’ pectin as a dual coagulant-flocculant
agent for heavy metals removal. Chemical Engineering Journal, 323, 19–28.
Jia, S., Yang, Z., Yang, W., Zhang, T., Zhang, S., Yang, X., Dong, Y., Wu, J. and Wang,
Y. (2016) Removal of Cu(II) and tetracycline using an aromatic rings-functionalized
chitosan-based flocculant: Enhanced interaction between the flocculant and the
antibiotic. Chemical Engineering Journal, 283, 495–503.
Khangembam, R., Tiwari, O. N. and Kalita, M. C. (2016) Production of
exopolysaccharides by the cyanobacterium Anabaena sp. BTA992 and application as
bioflocculants. Journal of Applied Biology & Biotechnology, 4(1), 8-11.
Klemm, D., Heublein, B., Fink, H.- P. and Bohn, A. (2005) Cellulose: Fascinating
biopolymer and sustainable raw material. Angewandte Chemie - International
Edition, 44(22), 3358–3393.
Kobielska, P. A., Howarth, A. J., Farha, O. K. and Nayak, S. (2018) Metal–organic
frameworks for heavy metal removal from water. Coordination Chemistry Reviews,
358, 92-107.
Kolya, H. and Tripathy, T. (2013) Preparation, investigation of metal ion removal and
flocculation performances of grafted hydroxyethyl starch. International Journal of
Biological Macromolecules, 62, 557-564.
Kono, H. (2017) Cationic flocculants derived from native cellulose: preparation,
biodegradability, and removal of dyes in aqueous solution. Resource-Efficient
Technologies, 3(1), 55-63.
Kono, H. and Kusumoto, R. (2015) Removal of anionic dyes in aqueous solution by
flocculation with cellulose ampholytes. Journal of Water Process, 7, 83-93.
Korish, M. (2015) Faba bean hulls as a potential source of pectin. Journal of Food
Science and Technology, 52(9), 6061–6066.
36 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
Liu, H., Yang, X., Zhang, Y., Zhu, H. and Yao, J. (2014a) Flocculation characteristics of
polyacrylamide grafted cellulose from Phyllostachys heterocycla: An efficient and
eco-friendly flocculant. Water Research, 59, 165–171.
Liu, J., Ma, J., Liu, Y., Yang, Y., Yue, D. and Wang, H. (2014b) Optimized production of
a novel bioflocculant M-C11 by Klebsiella sp. and its application in sludge
dewatering. Journal of Environmental Sciences, 26(10), 2076–2083.
Liu, J., Willför, S. and Xu, C. (2015) A review of bioactive plant polysaccharides:
Biological activities, functionalization, and biomedical applications. Bioactive
Carbohydrates and Dietary Fibre, 5(1), 31–61.
Liu, L., Jiang, T. and Yao, J. (2011) A Two-Step Chemical Process for the Extraction of
Cellulose Fiber and Pectin from Mulberry Branch Bark Efficiently. Journal of
Polymers and the Environment, 19(3), 568–573.
Liu, W. Zhao, C., Jiang, J., Lu, Q., Hao, Y., Wang, L. and Liu, C. (2015) Bioflocculant
production from untreated corn stover using Cellulosimicrobium cellulans L804
isolate and its application to harvesting microalgae. Biotechnology for Biofuels, 8(1),
1–12.
Liu, Z., Huang, M., Li, A. and Yang, H. (2017b) Flocculation and antimicrobial
properties of a cationized starch, 119, 57-66.
Liu, Z., Xu, D., Xia, N., Zhao, X., Kong, F., Wang, S. and Fatehi, P. (2018) Preparation
and application of phosphorylated xylan as a flocculant for cationic ethyl violet dye.
Polymers, 10(3), 1-16.
Lou, T., Cui, G., Xun, J., Wang, X., Feng, N. and Zhang, J. (2018) Synthesis of a
terpolymer based on chitosan and lignin as an effective flocculant for dye removal.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 537, 149-154.
Lou, T.,Wang, X., Song, G. and Cui, G. (2017) Synthesis and flocculation performance
of a chitosan-acrylamide-fulvic acid ternary copolymer. Carbohydrate Polymers,
170, 182–189.
Lu, X., Sun, W., Sun, Y. and Zheng, H. (2017) UV-initiated synthesis of a novel
chitosan-based flocculant with high flocculation efficiency for algal removal. Science
of the Total Environment, 609, 410–418.
Manivasagan, P., Kang, K., Kim, D. G. and Kim, S. (2015) Production of polysaccharide-
based bioflocculant for the synthesis of silver nanoparticles by Streptomyces sp.
International Journal of Biological Macromolecules, 77, 159-67.
Maran, J. P., Sivakumar, V., Thirugnanasambandham, K. and Sridhar, R. (2014)
Microwave assisted extraction of pectin from waste Citrullus lanatus fruit rinds.
Carbohydrate Polymers, 101(1), 786–791.
Marić, M., Grassino, A. N., Zhu, Z., Barba, F. J., Brnčić, M. and Brnčić, S. R. (2018) An
overview of the traditional and innovative approaches for pectin extraction from plant
food wastes and by-products: Ultrasound-, microwaves-, and enzyme-assisted
extraction. Trends in Food Science and Technology, 76, 28–37.
38 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
Mende, M., Schwarz, D., Steinbach, C., Boldt, R. and Schwarz, S. (2016) Simultaneous
adsorption of heavy metal ions and anions from aqueous solutions on chitosan—
investigated by spectrophotometry and SEM-EDX analysis. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 510, 275-282.
Mishra, A. and Bajpai, M. (2005) Flocculation behaviour of model textile wastewater
treated with a food grade polysaccharide. Journal of Hazardous Materials, 118(1–3),
213–217.
Mishra, A. and Bajpai, M. (2006) The flocculation performance of Tamarindus mucilage
in relation to removal of vat and direct dyes. Bioresource Technology, 97(8), 1055–
1059.
Mishra, A., Agarwal, M. and Yadav, A. (2003) Fenugreek mucilage as flocculating agent
for sewage treatment. Colloid and Polymer Science, 281(2), 164–167.
Mishra, A., Agarwal, M., Bajpai, M., Rajani, S. and Mishra, R. P. (2002) Plantago
psyllium Mucilage For Sewage and Tannery Effluent Treatment. Iranian Polymer
Journal, 1(16), 381–386.
Mishra, A., Bajpai, M., Pal, S., Agrawal, M. and Pandey, S. (2006) Tamarindus indica
mucilage and its acrylamide-grafted copolymer as flocculants for removal of dyes.
Colloid and Polymer Science, 285(2), 161–168.
Mishra, A., Srinivasan, R., Bajpai, M. and Dubey, R. (2004a) Use of polyacrylamide-
grafted Plantago psyllium mucilage as a flocculant for treatment of textile
wastewater. Colloid and Polymer Science, 282(7), 722–727.
Mishra, A., Yadav, A., Agarwal, M. and Bajpai, M. (2004b) Fenugreek mucilage for
solid removal from tannery effluent. Reactive and Functional Polymers, 59(1), 99–
104.
Mittal, H., Kumar, V., Alhassan, S. M. and Ray, S. S. (2018) Modification of gum ghatti
via grafting with acrylamide and analysis of its flocculation, adsorption, and
biodegradation properties. International Journal of Biological Macromolecules,
114(2017), 283-294.
Momeni, M. M., Kahforoushan, D., Abbasi, F. and Ghanbarian, S. (2018) Using
chitosan/CHPATC as coagulant to remove color and turbidity of industrial
wastewater: optimization through RSM design. Journal of Environmental
Management, 211, 347-355.
Moon, R. J., Martini, A., Nairn, J., Simonsenf, J. and Youngblood, J. (2011) Cellulose
nanomaterials review: structure, properties and nanocomposites. Chemical Society
Reviews, 40, 3941-3994.
Mukherjee, S., Mukhopadhyay, S., Pariatamby, A., Ali Hashim, M., Sahu, J. N. and Sen
Gupta, B. (2014) A comparative study of biopolymers and alum in the separation and
recovery of pulp fibres from paper mill effluent by flocculation. Journal of
Environmental Sciences, 26(9), 1851-1860.
Flocculant Polysacharides Mainly from Plants 39
Müller-Maatsch, J., Bencivenni, M., Caligiani, A., Tedeschi, T., Bruggeman, G., Bosch,
M., Petrusan, J., Droogenbroeck, B. V., Elst, K. and Sforza, S. (2016) Pectin content
and composition from different food waste streams. Food Chemistry, 201, 37–45.
Okaiyeto, K., Nwodo, U. U., Okoli, S. A., Mabinya, L. V. and Okoh, A.I. (2016)
Implications for public health demands alternatives to inorganic and synthetic
flocculants: Bioflocculants as important candidates. Microbiology Open, 5(2), 177–
211.
Oladoja, N. A., Unuabonah, E. I., Amuda, O. S. and Kolawole, O.M. (2017) Mechanistic
insight into the coagulation efficiency of polysaccharide-based coagulants. In:
Polysaccharides as a green and sustainable resources for water and wastewater
treatment. Biobased Polymers, Springer International Publishing, 13-35.
Oliveira, A. N., Paula, D. A., Oliveira, E. B., Saraiva, S. H., Stringheta, P. C. and Ramos,
A. M. (2018) Optimization of pectin extraction from Ubá mango peel through surface
response methodology. International Journal of Biological Macromolecules, 113,
395–402.
Oliveira, T. Í., Rosa, M.F., Cavalcante, F. L., Pereira, P. H., Moates, G. K., Wellner, N.,
Mazzetto, S. E., Waldron, K. W. and Azeredo, H. M. (2016) Optimization of pectin
extraction from banana peels with citric acid by using response surface methodology.
Food Chemistry, 198, 113–118.
Ortiz, J. A., Matsuhiro, B., Zapata, P. A., Corrales, T. and Catalina, F. (2018) Preparation
and characterization of maleoylagarose/PNIPAAm graft copolymers and formation
of polyelectrolyte complexes with chitosan. Carbohydrate Polymers, 182, 81–91.
Pal, S., Mal, D. and Singh, R. P. (2005) Cationic starch: An effective flocculating agent.
Carbohydrate Polymers, 59(4), 417–423.
Pal, S., Patra, A. S., Ghorai, S., Sarkar, A. K., Das, R. and Sarkar, S. (2015) Modified
guar gum/SiO2 : development and application of a novel hybrid nanocomposite as a
flocculant for the treatment of wastewater. Environmental Science: Water Research
Technology, 1(1), 84-95.
Patel, M., Patel, U. and Gupte, S. (2014) Production of exopolysaccharide (EPS) and its
application by new fungal isolates SGMP1 and SGMP2. International Journal of
Agriculture, Environment & Biotechnology, 7, 511-523.
Pereira, P. H. F., Oliveira, T. Í. S., Rosa, M. F., Cavalcante, F. L., Moates, G. K.,
Wellner, N., Waldron, K. W. and Azeredo, H. M. C. (2016) Pectin extraction from
pomegranate peels with citric acid. International Journal of Biological
Macromolecules, 88, 373–379.
Pérez, S. and Bertoft, E. (2010) The molecular structures of starch components and their
contribution to the architecture of starch granules: A comprehensive review.
Starch/Staerke, 62(8), 389–420.
40 P. B. S. Albuquerque, W. F. Oliveira, P. M. dos Santos Silva et al.
Subudhi, S., Bisht, V., Batta, N., Pathak, M., Devi, A. and Lal, B. (2016) Purification and
characterization of exopolysaccharide bioflocculant produced by heavy metal
resistant Achromobacter xylosoxidans. Carbohydrate Polymers, 137, 441-451.
Sun, Y., Zhu, C., Sun, W., Xu, Y., Xiao, X., Zheng, H., Wu, H. and Liu C. (2017)
Plasma-initiated polymerization of chitosan-based CS-g-P(AM-DMDAAC)
flocculant for the enhanced flocculation of low-algal-turbidity water. Carbohydrate
Polymers, 164, 222–232.
Suopajärvi, T., Liimatainen, H., Hormi, O. and Niinimäki, J. (2013) Coagulation–
flocculation treatment of municipal wastewater based on anionized nanocelluloses.
Chemical Engineering Journal, 231, 59–67.
Szyguła, A., Guibal, E., Palacín, M. A., Ruiz, M. and Sastre, A. M. (2009) Removal of an
anionic dye (Acid Blue 92) by coagulation-flocculation using chitosan. Journal of
Environmental Management, 90(10), 2979-2986.
Tamnak, S., Mirhosseini, H., Tan, C. P., Ghazali, H. M. and Muhammad, K. (2016)
Physicochemical properties, rheological behavior and morphology of pectin-pea
protein isolate mixtures and conjugates in aqueous system and oil in water emulsion.
Food Hydrocolloids, 56, 405–416.
Tang, J., Qi, S., Li, Z., Na, Q., Qu, M., Xie, M., Yang, B. and Wang, Y. (2015)
Evaluation of metal ions on the production and flocculating activity of
polysaccharide-based bioflocculant from Paenibacillus mucilaginosus.
Biotechnology: An Indian Journal, 11(7), 255-264.
Teh, C. Y., Budiman, P. M., Shak, K. P. Y. and Wu, T. Y. (2016) Recent advancement of
coagulation-flocculation and its application in wastewater treatment. Industrial and
Engineering Chemistry Research, 55(16), 4363-4389.
Teh, C. Y., Wu, T. Y. and Juan, J. C. (2014) Potential use of rice starch in coagulation-
flocculation process of agro-industrial wastewater: treatment performance and flocs
characterization. Ecological Engineering, 71, 509-519.
Toczyłowska-Mamińska, R. (2017) Limits and perspectives of pulp and paper industry
wastewater treatment – a review. Renewable and Sustainable Energy Reviews, 78,
764-772.
Vakili, M., Rafatullah, M., Salamatinia, B., Abdullah, A. Z., Ibrahim, M. H., Tan, K. B.,
Gholami, Z. and Amouzgar, P. (2014) Application of chitosan and its derivatives as
adsorbents for dye removal from water and wastewater: a review. Carbohydrate
Polymers, 113, 115-130.
Verma, A. K., Dash, R. R. and Bhunia, P. (2012) A review on chemical
coagulation/flocculation technologies for removal of colour from textile wastewaters.
Journal of Environmental Management, 93(1), 154-168.
Wang, D., Zhao, T., Yan, L., Mi, Z., Gu, Q. and Zhang, Y. (2016) Synthesis,
characterization and evaluation of dewatering properties of chitosan-grafting
Flocculant Polysacharides Mainly from Plants 43
BIOGRAPHICAL SKETCH
Education: PhD
1. https://www.sciencedirect.com/science/article/pii/S0141813016324394
2. https://www.ncbi.nlm.nih.gov/pubmed/28433769
3. https://www.ncbi.nlm.nih.gov/pubmed/28916381
4. https://www.ncbi.nlm.nih.gov/pubmed/28987799
5. https://www.ncbi.nlm.nih.gov/pubmed/28545372
6. https://www.ncbi.nlm.nih.gov/pubmed/26840177
7. https://www.ncbi.nlm.nih.gov/pubmed/29632933
8. https://www.ncbi.nlm.nih.gov/pubmed/26428171
9. http://www.journals.ufrpe.br/index.php/JEAP/article/view/1701
10. http://www.sciencedomain.org/abstract/13497
11. http://www.aimspress.com/article/10.3934/molsci.2016.3.386
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 2
ABSTRACT
INTRODUCTION
The coagulation and flocculation processes are decisive in the treatment of water,
both in purification and in wastewater, due to their efficiency of removing suspended
particles from the water. To carry out the coagulation, it is necessary to add substances
called coagulants, which come from chemical reactions that have been industrialized.
The most important task is to generate drinking water, which can help the inhabitants
of underdeveloped countries to improve their living conditions, as it is established in the
main policies of the United Nations; but reducing the environmental impact of these
activities and improving the coagulation and flocculation processes with coagulants of
organic origin are easy to produce in agriculture and without metals that can affect the
ecosystem from the ground.
This document shows in a general way the advances in the processes of coagulation
and flocculation with plants, and an emphasis is made on showing those made in recent
years.
The presented study is based on experiences with biopolymers derived from plants,
especially from Melocactus sp, Opuntia sp, Stenocereus griseus, Cereus forbesii, Aloe
arborescens, Aloe vera and Cicer arietinum. Melocactus sp, Stenocereus griseus, and
Cereus forbesii have not been studied very much by other authors.
Conventional water treatment processes include coagulation and flocculation,
especially that which is chemically assisted, for both drinking water and wastewater; the
removal capacity of these operations are effective and possess a speed that could have
been used, especially with waters that have high concentrations of organic matter, and for
this purpose reagents have been used such as aluminum sulfate, which is usually obtained
from the reaction of aluminum hydroxide with sulfuric acid. Another reagent used is
ferric chloride, which is obtained via the reaction of chlorine gas on heated iron, and an
additional reagent that has been used in recent years is the aluminum polychloride that is
obtained through a reaction of aluminum with hydrochloric acid in an aqueous solution.
The reactions associated with the coagulants are determined by some parameters such
as the pH, the temperature of the water, and the concentrations of the solids that are going
to be complexed to form flocs that can be separated by density difference inside the
mixture. When the flocs are separated from the mixture, sludge is generated that must be
thickened and then disposed of within the parameters of a waste management plan. These
residues will have a high concentration of aluminum and iron, respectively, according to
the type of coagulant used, whether it be aluminum sulfate, aluminum polychloride or
ferric chloride. The final disposal of this sludge usually comes with some difficulties,
Coagulation and Flocculation with Plant Extracts 49
because due to their load of aluminum or iron they are considered toxic for soil in high
concentrations.
The disposal problems of this sludge, which generally has a high concentration of
organic matter, generates environmental impacts when they are discharged into soils or
bodies of water, changing the natural microbiota and affecting the species that have
contact with these high concentrations of aluminum and iron. Because of this, different
products of vegetable origin have been studied, which have properties similar to those of
aluminum or iron compounds and generate coagulation and flocculation with organic
compounds. Some examples of these cases include Melocactus sp, Opuntia dilleni,
Stenocereus griseus, Cereus forbesii, Aloe arborescens, Aloe vera and the Kabuli
Chickpea (Cicer arietinum L); these plants have shown activity for the flocculation of
substances with small particle sizes (below 0.2 mm), which generally cannot be separated
by natural sedimentation.
The sludge derived from the coagulation and flocculation processes with plant
extracts has a completely organic composition, which means that they can be digested by
microorganisms and transformed into carbon, nitrogen and phosphorus substances that
can be incorporated into the corresponding biogeochemical cycles, along with the
absence of toxic metals for the soil, or with safe concentrations for this vital resource; this
technological alternative transforms water treatment into a less aggressive process with
the environment, taking into account that most of the waste generated in drinking water
and domestic waste treatment is sludge.
The extraction systems of plant biopolymers have different methodologies, which are
easy to apply, proven and are part of already standardized unit operations. Taking into
account that different parts are harvested from each plant, we must understand that for
most of the plants their use concerns the majority of the biomass, whereas when we speak
of Cicer arietinum, we are using only their seeds. This diminishes its use, taking into
account the weight ratio of the plant and the mass used for the preparation of the
coagulant.
The operations developed to determine the efficiency of each plant extract in the
coagulation and flocculation operations are defined within the established jar tests, and
some of them have Z potential measurements (a measure of the magnitude of the
repulsion or attraction between the particles).
CONVENTIONAL COAGULANTS
Coagulants are substances with the capacity to break the stability of aqueous
mixtures, causing suspended particles to be grouped in flocs, to be later segregated by
gravity or by air injection separators.
50 Jesús Manuel Epalza Contreras and Johan Jaramillo Peralta
Aluminum sulfate [Al2 (SO4) 3.x18 H2O]: This substance is an inorganic salt
resulting from the controlled mixing of sulfuric acid and aluminum or bauxite,
according to the purity and concentration desired for the final product.
Ferric chloride [FeCl3. x6 H2O]: It is an inorganic salt, which is a product of the
reaction between metallic iron and dilute commercial hydrochloric acid; this
reaction is generally carried out with the industrial leftovers of hydrochloric acid,
resulting in a high concentration ferric chloride.
Aluminum Chloride [Al2 (OH) 3Cl or Al2 (OH) 3Cl3]: This salt is the result of
the reaction of aluminum, bauxite or other minerals with high concentrations of
aluminum and hydrochloric acid; from which different types of aluminum
polychloride are produced, such as those that are named as PAC or PAFC,
depending on their specification. Their alumina content may be between 13% and
17%.
The use of these salts has an environmental impact, especially regarding the
generation of large quantities of sewage sludge with high concentrations of metals,
usually either aluminum or iron.
The final disposal of this sewage sludge is complicated, as their high aluminum
content makes them toxic to the soil and plants that will grow in this site (Casierra Posada
Fanor, 2007).
Less commonly, other coagulating substances may be used, such as:
to extract the coagulant biopolymer, preferably in a similar way to the one that performs
with conventional coagulants.
The plants studied for the raw material of this process are diverse; within them, we
have Moringa oleifera, Opuntia spp, Jatropha curcas, Cicer arietinum, Melocactus spp,
Stenocereus griseus, Cereus forbesii, Aloe arborescens, Aloe vera, Cicer arietinum and
Coccinia indica; these plants have presented good performance to coagulate and
flocculate raw water, leading to better sanitary conditions.
In the case of this text, we explored the species Melocactus sp, Opuntia sp,
Stenocereus griseus, Brain forbesii, Aloe arborescens, Aloe vera and Cicer arietinum L,
which were tested in their efficiency for coagulation and flocculation with raw water.
Plants with the capacity to generate biopolymers with coagulants and flocculants
have been studied over the last few decades, especially Moringa olerifera, Opuntia spp,
Cicer arietinum, and others that have demonstrated coagulant capacity as part of the
traditional empirical knowledge of indigenous communities.
Source: http://cactiguide.com/cactus/?genus=Stenocereus&species=griseus.
The selected plants took coincided with those referenced and others present in semi-
arid regions in Colombia, such as La Guajira in northern Colombia and the banks of the
Chicamocha River in the northeastern region; the species not studied are Stenocereus
griseus (Figure 1), Cereus forbesii (Figure 2), Aloe arborescens, and Aloe vera; and the
one already studied was Cicer arietinum (Moa Megersa, 2014).
In the case of Opintia sp, they are present in the Colombian regions already named
and sufficiently studied, in the same way Aloe vera and Aloe arborecens were selected
52 Jesús Manuel Epalza Contreras and Johan Jaramillo Peralta
(Figure 3); the Kabuli Chickpea (Cicer arietinum) is reviewed with studies
already elaborated on by other authors (Hildebrando Ramírez Arcila, 2015). It
should be noted that some species of Opuntia spp are used as part of the animal and
human diet in communities of northeastern Colombia in semi-desert areas
(Fernández, 2002).
Source: http://www.kakteensammlungholzheu.de/en/cereus_forbesii.html.
To have a better selection, we reviewed the massive presence of these plants and they
were not part of the list of plants in danger of extinction.
The extractions of each plant have particularities, taking into account the usable parts
in the search of their coagulating capacity; the extractions are segregated into two types:
The plants that are used in all their foliage are Melocactus sp (Figure 6), Opuntia dilleni
Coagulation and Flocculation with Plant Extracts 53
(Figure 4), Stenocereus griseus, and Cereus forbesii, which belong to the family
Cactaceae while the other part corresponds to Aloe arborescens, and Aloe vera, which
belongs to the Xanthorrhoeaceae family, with superior Aloe classification.
For the cactus and aloe species, an extraction methodology was used with different
operations that are:
The selection of parts of the plant for cutting: The operation of cutting parts of
the plant is done by taking into account the mature parts, with the presence of
thorns in the case of cacti and a hard external surface, similar to the criteria used
for the animal or human consumption of Opuntia spp species. In the case of the
Aloe species, the maturity of the leaves is considered, with the presence of
perimeter spines, as this shows the possibility of isolating the crystals of the
plant. Figure 4 shows a part of the penca or cladodes of a catus Opuntia sp with
skin and thorns, but the part to be used is the vascular tissue of the plant,
eliminating the skin and spines. For the sampling of the species of Aloe (Figure
5) garden plants were considered, which are cultivated in a homemade way,
taking into account the age of the plant, as it must have enough leaves with
enough crystals, and it must not present any evidence of contamination or
parasites, especially the characteristics of the green color of the leaves, absence
of external insects and total absence of organisms associated with diseases of the
plant.
Cut and transport to the laboratory. The cutting of the parts of the plant takes
measurements (Figure 7) for transportation to the site of the coagulant; the
transport must be carried out in refrigeration to avoid possible contamination
with environmental fungi or other organisms that can significantly change the
composition of the parts obtained.
Source: http://www.fichas.suculentas.es/Almacenfichas/903/903.html
Source: http://www.fichas.suculentas.es/Almacenfichas/903/903.html.
Weighing of the gross material. The weighing of the material is carried out on a
25 kg scale to determine the weight of the sample taken and then determine its
performance according to its humidity.
Cutting of thorns and removal of the bark. To perform the extraction of the
coagulant, the cut parts are taken and the thorns are removed (Figure 8) along
with the skin of the cacti, also called the epidermis, which is the external hard
part of the pads or cladodes in the case of cacti and leaves in the case of Aloe
species.
Source: http://www.tephroweb.ch/kuas/melo.htm.
Cutting of clean material. In order to follow the extraction process, the tissue of
the plants is cut; this is done to improve the loss of moisture in the plant,
increasing the contact surface with the atmosphere.
Drying the pieces. The drying of the material is carried out outdoors (Figure 10),
and it is important to note that the region in which these operations are carried
out are from a warm tropical climate with low humidity. In these areas, the
Coagulation and Flocculation with Plant Extracts 55
temperature ranges between 20°C and 35°C with relative humidity between 50%
and 70% on average.
Source: authors.
Source: authors.
Source: authors.
Weighing of dehydrated material. The material is kept outdoors, taking care that
it is not hydrated by rain, and the drying time is between 48 and 96 hours
depending on the temperature and relative humidity of the place where they are
dried.
Grinding of the material. The grinding of the dried material is done with a food
processing device (Photo 1) and then it is passed through a mill that pulverizes
the material. The final characteristics of the material are similar to the raw
materials used in the coagulation and flocculation process, that is, a presentation
similar to the presentation of type A and B aluminum sulfate and aluminum
polychloride in their solid presentations. The liquid presentation is not sought
because it is an organic material with nutritional characteristics for filamentous
fungi and bacteria, which would require a procedure for its sterilization. For any
application of excessive heat, a degradation of the biopolymers will be carried
out, which is a condition that impairs its performance in coagulation.
Material screening. The material after being macerated is taken to a 1 mm sieve
(Figure 11), which separates the thick parts that cannot be diluted efficiently in
an aqueous solution; this characteristic is based on the solubility of complex
organic substances, such as the vascular tissue of these cacti and Aloe plants.
Weighing of the material of interest. The material obtained from the plants
Melocactus sp, Opuntia dilleni, Stenocereus griseus, Cereus forbesii, Aloe
arborescens and Aloe vera is weighed to have a reference of its performance in
relation to the weight, with respect to its use.
Source: authors.
The seeds of the anionic coagulant Kabuli Garbanzo (Cicer arietinum L) were
selected with the following procedures being performed.
Washed. The seeds selected without evidence of the presence of fungi or yeasts were
washed with large amounts of water to eliminate impurities related to bulk handling.
In terms of their fractionation and packaging, they can take other materials such as
small sand stones or other grain waste.
Drying. After washing, they were dried for two days in the sun, taking into account
that they cannot be wetted by rain or other water sources.
Crushed. The material was crushed using a mixer (Oster) and the resulting powder
was sieved with a No. 200 sieve to obtain a very fine powder for storage in plastic
containers to avoid hydration and subsequent use in the preparation of the solutions
of the coagulant.
After obtaining the biopolymers of the plants Melocactus sp, Opuntia dilleni,
Stenocereus griseus, Cereus forbesii, Aloe arborescens and Aloe vera, the preparation of
the solutions is carried out to realize the tests of the jars, which prove the action of the
coagulants and flocculants.
The flocculant preparation process was weighed for 1 gram of a biopolymer; then,
1000 ml of distilled water was added and manual agitation was carried out until it was
completely diluted.
To test the coagulating and flocculating effect of the biopolymers of Melocactus sp,
Opuntia dilleni, Stenocereus griseus, Cereus forbesii, Aloe arborescens, Aloe vera and
the Kabuli Chickpea (Cicer arietinum L), jug tests were performed as provided in ASTM
D2035: 08
The control parameters normally used in the efficiency of a coagulant are pH,
turbidity and color, which are governed by standardized methodologies and which
determine the ability to remove solids from water based on their behavior. pH and its
spectrophotometric absorbance in the case of turbidity and color at different wavelengths
were studied.
Coagulation and Flocculation with Plant Extracts 59
In accordance with the regulations in force in Colombia, the tests were carried out
taking into account the technical standards adopted by the country for each test.
The removal obtained is processed in terms of percentages for the comparison of pH.
It is observed that it complies with the provisions of the national standard for this
parameter; and in the case of turbidity and color, the percentage of removal is taken into
account, which is directly proportional to the decrease in absorbance in each test.
For the results of the tests of the biopolymers, natural coagulants are collected in 4
main tests: The first ones refer to the pH, the turbidity and the color, and the last one
takes the coagulants with good performance and measures the Z potential.
Turbidity and color in water is related to the presence of substances or
microorganisms, which is directly related to its quality to be consumed or used in other
ways; the results offer an overview of the potentialities of all the plants for the
coagulation and flocculation processes, with different degrees of efficiency.
To control the efficiency of coagulation and flocculation as a function of pH, the pH
was taken after coagulation, taking into account that the initial pH of the water is 7.2. The
pH results are shown in Graph 1, showing that the coagulant that most affected the final
pH was the biopolymer of Melocactus sp, which brought the pH up to 6.2. The data show
a standard deviation of 0.3
Graph 1 shows that the only one that did not affect the pH in the jar test was
Stenocereus spp, while the others lowered the pH moderately to values of 7 or 6.8.
The turbidity results (Graph 2) showed that the best biopolymers to remove these
solids associated with turbidity were Melocactus spp and Cicer arietinum, which showed
turbidity removals greater than 95% and 97% respectively. The data show a standard
deviation of 3.1.
The results of the other biopolymers showed a removal capacity greater than 88%
and up to 92%, which shows the effectiveness of these biopolymers with water that has a
neutral pH.
All the biopolymers tested showed effective action in the removal of turbidity, in a
range between 88% and 97%, with some differences and affectations to the pH of the
sample at the final moment of the jar test.
For the case of the color results (Graph 3), we can see a good activity of all the
biopolymers, taking into account that the one that showed the best performance in the
removal of the color was the biopolymer of Melocactus spp, with a performance greater
than 96%, which was then appreciated for the two species of Aloe sp with a removal
greater than 95%. Opuntia sp, Stenocereus sp and the Kabuli Chickpea (Cicer arietinum
L) displayed performances greater than 94%, and the last one was evidenced
60 Jesús Manuel Epalza Contreras and Johan Jaramillo Peralta
by Cereus forbesi with a performance greater than 92%. The data show a standard
deviation of 1.4.
Source: authors.
Source: authors.
Source: authors.
The results show a greater efficiency in the removal of color; the best
performance of these tests lies in the Melocactus, which was the best in removing
both turbidity and color, but also had the highest incidence of pH. The
other biopolymer with a high performance is the Kabuli Chickpea (Cicer
arietinum L), whichshowed a good turbidity and color removal capacity with very little
pH affectation.
The zeta potential of the biopolymers shows similar values in the range of pH 4 and
pH 10, which may indicate a similar activity with solid particles of a small size, such as
those that generate turbidity and color in the water.
62 Jesús Manuel Epalza Contreras and Johan Jaramillo Peralta
The zeta potential measurements were made for the biopolymers with better
performance, which showed the following results.
Table 1 shows the results of the Z potential of Melocactus sp and the Kabuli
Chickpea (Cicer arietinum L), with a pH between 3 and 10.
It is important to note that all plants used have the potential to treat water; the
efficiency differences can be associated with the affinity for different particles and their
extraction form.
CONCLUSION
All the extracts showed turbidity and color removal with efficiencies higher than
88%, which indicates that the extraction methodologies conserve the coagulant and
flocculant capacity of each plant.
The biopolymers of Melocactus sp, Opuntia sp, Stenocereus griseus, Cereus forbesii,
Aloe arborescens, Aloe vera and Cicer arietinum have an activity for coagulation and
water flocculation.
The plants with the best performance in the removal of turbidity and color were
Melocactus sp and Cicer arietinum, with the best percentages of elimination for solids of
small size in water.
The potential zeta measurements for the extracts of Melocactus sp and Cicer
arietinum have similar values in the range of pH 4 to 10, which shows a similar activity
for the suspended particles of the water used in the tests.
The biopolymers of the plants Melocactus sp, Opuntia sp, Stenocereus
griseus, Cereus forbesii, Aloe arborescens, Aloe vera and Cicer arietinum can be
a viable alternative for the treatment of drinking and residual water, in terms
of the replacement of sulfate for aluminum, aluminum polychloride and
ferric chloride; this would allow for decreasing the amounts of dissolved metals in the
drinking water of humans and animals, especially with the aluminum residues associated
with diseases such as autism and Alzheimer’s.
ACKNOWLEDGMENTS
CONFLICT OF INTEREST
The authors declare that there is no conflict of interest regarding the publication of
this document.
REFERENCES
BIOGRAPHICAL SKETCH
Chapter 3
ABSTRACT
The present study aimed to use the organic polymer from Opuntia cochenillifera
cactus associated with the addtion of aluminum sulfate to treat the water of a lentic body
(Igapó II Lake - Located in Londrina in the state of Paraná and Brazil) applying
coagulation, flocculation, sedimentation and filtration processes. The tests were
performed with static a Jar-test reactor, aiming to evaluate the removal of parameters of
turbidity and of apparent color, besides monitoring the pH variation. Regarding the
physico-chemical characteristics, there was little influence of the Opuntia cochenillifera
cactus in the pH parameter tending to neutrality. Therefore, the study showed that the use
of an organic coagulant together with aluminum sulphate is effective for the removal of
parameters of apparent color and of turbidity (95 and 94%, respectively).
1. INTRODUCTION
Water intended for human consumption must be available in a potable way, in which
it does not present health risks and denote the control and monitoring of its quality,
according to the drinking standard established by the Ministry of Health Ordinance No.
2,914/2011 (Brazil 2011).
In order to reach the standards established through the water regulatory agencies and
to distribute it to a community, it is necessary to carry out its treatment (Howe et al.
2016). In a conventional manner, the water treatment consists of the following steps,
which are shown in Figure 1.
The clarification of the raw water occurs in the removal of solids present and
associated with the parameter of turbidity, being removed by operations and processes as:
coagulation, flocculation, sedimentation and filtration. At the beginning of the
coagulation process, the coagulant is added under intense stirring in the water to be
treated, ensuring uniform distribution of the product. This phase has the objective of
destabilizing the impurities present in the medium, with the addition of a coagulant,
capable of minimizing or eliminating the repulsive forces present between them. The
most commonly used coagulants are metal salts based on aluminum or iron (COMUSA
2017; SANEP 2018). In order to replace or reduce the use of conventional coagulants, the
natural agents with equal or superior efficiency when compared to metallic ones,
generating a sludge with less toxicity, have a lower cost and greater abundance of plants
as a source of extraction available (Yin 2010; Theodoro et al. 2013).
Derived from a renewable resource and its arrangement promotes simple organic
degradation, mucilage of cacti for extraction of polymers is used to eliminate water
turbidity (Pichler; Alcantar, 2012).
The Process of Water Treatment with Aluminum Sulphate ... 71
The forage cactus Opuntia cochenillifera, originating in Mexico and spread in the
northeast of Brazil, has several uses, ranging from animal and human feeding, to
landscaping projects and dye extraction. Among its characteristics, it has the xeromorphic
form with a cylindrical stem and its branches known as palms.
Thus, the objective of this chapter was to analyze the water quality of Lake Igapó II,
found in the city of Londrina-PR, according to the parameters of apparent color, of
turbidity and of pH, and then to perform the treatment of these waters through the
coagulation/flocculation/ sedimentation/filtration process using the organic polymer from
the Opuntia cochenillifera cactus associated with the addition of Aluminum Sulphate,
verifying the possibility use of these lentic waters for urban supply.
2. METHODOLOGY
The water samples were taken from a lentic system from Lake Igapó II, located in the
city of Londrina, state of Paraná, Brazil.
Situated at 550 meters above sea level, one has the coordinates of the collection point
are 23°19'42'' S and 51°10'11'' W. Figure 2 spatially represents the location of the
highlighted point for sample collection. To perform the tests, 72 liters of Lake Igapó II
water were collected and stored in previously sanitized plastic gallons.
The methodology for the extraction of the polymer present in the cactus occurred in
two stages: The first consisted in the removal of the spines and the bark, and thus cuts
were made to reduce the size of the cactus to be used in the liquidification of the contents
inside the palms ("pulp").
In the second, two liters of the extraction solution (required to extract the polymer
from the viscous solution formed after liquefying) were prepared by the attachment of
one liter of distilled water with 4g of 1% sodium chloride solutions and one liter of
distilled water with 10g sodium hydroxide 0.10mol L-1 (Zara et al. 2012).
The ratio was one ml of cactus prepared to 2.5 mL of the extraction solution, the
mixture of which was homogenized on magnetic stirrer Nova Ethics, model 114, for 40
minutes. The resulting viscous complex was packed in glass sanitized vials and stored
under refrigeration at 5°C until assays were performed on the JarTest equipment (Goes et
al. 2017). Some of these steps are shown in Figure 3.
The amount of water collected was divided into three gallons with the same volume,
to carry out the tests at different pH values, being basic, neutral and acidic sample in
order to work with the parameters of color and of turbidity after the operations and
processes in the three tracks. To adjust the pH values with the purpose of making it
acidic, the hydrochloric acid solution (HCl) with1 M concentration was used.
For the basic pH sample, the sodium hydroxide solution (NaOH) with 1 M
concentration. The neutral pH value of the crude sample was also measured.
For the experiment, the dosages of coagulant cacti (Opuntia cochenillifera) were 1mg
L , 4mg L-1 and 7mg L-1 with the addition of Aluminum Sulphate in 1mg L-1, 4mg L-1
-1
Figura 3. a) Opuntia cochenillifera Palm; b) Removing the shell; c) Cut for further liquefaction; d)
Storage of viscous complex.
The Process of Water Treatment with Aluminum Sulphate ... 73
Quick Mix Slow Mix Slow Mix Slow Mix Slow Mix
1 2 3 4
Gradient (s-1) Gradient (s-1) Gradient (s-1) Gradient (s-1) Gradient (s-1)
450 90 52 40 30
Time (mim) Time (mim) Time (mim) Time (mim) Time (mim)
00:10 02:00 02:35 02:40 05:40
Source: Higashi (2016).
The tubes are 25cm long with 15cm filled by sand, 3cm filled with crushed
stone and 2cm filled with cotton. The granulometric of the sand used in the 6 filters
were the same, in the range of 0.60 to 0.85mm. The slope used at the exit of the
jar-test was 70º for two minutes, followed by another two minutes with the
angulation of 60º and another two minutes with 50º (Di Bernardo et al. 2003). The water
to be treated entered the top of the column and its supernatant was removed from the
side.
The grains of the filter bed were retained by a coffee filter together with the cotton
and stone, present at the lower end of the column.
For all variations of cactus concentrations and pH, the parameters of apparent color
and of turbidity were determined according to the Standard Methods of Examination of
Water and Wastewater (APHA 2012). In this work we used a statistical planning with
two factors (independent variables), the concentration of coagulant cacti and three
parameters responses (apparent color, turbidity and pH), the tests were performed in
duplicates.
3. RESULTS
The collected raw water had a neutral pH of 7.65. For acid pH and base, we have the
values obtained of 4.87 and 9.08 respectively.
After the treatment, the calculations were performed to remove the parameters
apparent color and turbidity, thus obtaining data after the coagulation/flocculation/
sedimentation operation/filtration process and the overall process/global operation
(coagulation/flocculation/sedimentation/ filtration), besides the monitoring of pH. The
values found are shown in Table 2.
Figures 5, 6 and 7 show the removal of the parameter apparent color and
turbidity for the coagulation/flocculation/sedimentation processes, only filtration
and for the complete process (coagulation/flocculation/ sedimentation/filtration),
respectively.
Analyzing Figure 5, it can be seen that the best results presented are associated with
the neutral pH values (tests 2, 5, 8, 11, 14 and 17), removing around 30% of the color and
turbidity. It is possible to verify that in tests 1, 3, 4, 6, 9, 10, 12, 15 and 18 the dissolution
of the organic coagulant cactus occurred in the solution, causing the increase of the
dissolved solids and, thus, and, thus, justify the negative values of percentage of removal
of the color parameter.
Table 2. Numerical organization of the tests for different levels of coagulant concentration and pH with real values and
results of parameters color, turbidity and pH
The filtration process was more efficient for the basic and acid pH values, allowing a
removal of more than 80% in the two analyzed parameters. With the completion of the
complete process, it was noticed that in general the tests obtained values of color removal
and turbidity above 80%. Thus, by analyzing the presented results, it can be concluded
that both the coagulation/flocculation/sedimentation processes (Figure 5) and the
filtration process (Figure 6) are important for the removal of color and of turbidity
present in the raw water. Comparing the values obtained with those required by
Ordinance 2,914/2011 of the Ministry of Health (Brazil 2011), the tests did not meet the
VMP (maximum allowed value) of apparent color and of turbidity. The results of pH
The Process of Water Treatment with Aluminum Sulphate ... 77
After also passing through the filtration (Figure 9), the water tends to obtain values
closer to the neutrality.
This occurs due to the ability of the sand particles to adsorb the components present
in the water, neutralizing the pH. Comparing the values obtained with those required by
78 H. H. Dias Goes, R. de C. P. de Souza and J. D. Peruço Theodoro
Portaria 2,914 / 2011 (Brazil 2011) of the Ministry of Health, all the tests met the pH
VMP (6 to 9).
CONCLUSION
According to the tests carried out in the study, the coagulant solution obtained from
the Opuntia cochenillifera cactus associated with the addition of aluminum sulphate was
efficient in the treatment of water from the Lake Igapó lentic system, tending to maintain
its pH close to neutrality. The results for color removal and turbidity were promising, but
did not reach compliance with the values required by the Ministry of Health Ordinance
2,914/2011 (Brazil 2011), not meeting the VMP (maximum value allowed) for both
parameters.
REFERENCES
APHA – American Public Health Association. Standard Methods for the Examination of
Water and Wastewater. 22 ed. Washington, 2012.
Brazil, Ministry of Health. 2,914, dated December 12, 2011. Available at:
<http://bvsms.saude.gov.br/bvs/saudelegis/gm/2011/prt2914_12_12_2011.html>.
Accessed on: January, 2018.
COMUSA, Water and Sewage Service of Novo Hamburgo. Water treatment, 2017.
Available at: http://www.comusa.rs.gov.br/index.php/saneamento/tratamentoagua.
Accessed on: Feb. 2017.
The Process of Water Treatment with Aluminum Sulphate ... 79
Di Bernardo, L.; Mendes, C. G. N.; Brandão, C. C. S.; Sens, M. L.; Pádua, V. L. Water
Treatment for Direct Filtration. Di Bernardo, Luiz (coordinator) - Rio de Janeiro:
ABES, RiMa, 2003. Project PROSAB 498 p.
Goes, H. H. D.; Souza, R. C. P.; Melo, J. M.; Theodoro, J. D. P.; Study of the application
of Opuntia cochenillifera cactus in water treatment. Encyclopedia Biosphere,
Knowing Scientific Center - Goiânia, v.14 n.25; P. 554-563. 2017.
Higashi, V. Y.; Theodoro, J. D. P.; Pereira, E. R.; Theodoro, P. S.; Use Of Chemical
Coagulants (Ferric Chloride) And Organic (Moringa Oleifera) In Treatment Of
Waters Derived From The Lentic System. Technical Scientific Congress of
Engineering and Agronomy - CONTECC'2016, August 29 to September 2, 2016 -
Foz do Iguaçu, Brazil. Available at: <https://www.tratamentodeagua.com.br/wp-
content/uploads/2016/11/Uso-de-coagulantes-qu%C3%ADmico-cloreto-ferrico-e-
org%C3%A2nico-moringa-oleifera-in-treatment-of-water-from-system-l%C3%
AAntico.pdf>. Accessed in Feb. 2018.
Howe, K. J.; Hand, D. W.; Crittenden, J. C.; Trussell, R. R.; Tchobanoglous, G.;
Principles of water treatment, São Paulo, SP: Cengrage, 2016. 624 p.
Pichler, T., Young, K.; Alcantar, N. Eliminating turbidity in drinking water using the
mucilage of a common cactus. Water Science and Technology: Water Supply, v. 12,
n.2, 179-186, 2012.
SANEP, Autonomous Service of saniamento of Pelotas. Treatment, 2018. Available at:
<http://server.pelotas.com.br/sanep/tratamento/>. Accessed on: Feb. 2018.
Theodoro, J. D. P.; Lenz, G. F.; Zara, R. F.; Bergamasco, R.. Coagulants and Natural
Polymers: Perspectives for the Treatment of Water. Plastic and Polymer Technology
(PAPT), v. 2, Issue 3, September 2013.
Yin, C. Emerging usage of plant0based coagulants for water na wastewater treatment.
Process Biochemistry, v. 45, 2010.
Zara, R. F.; Thomazini, M. H.; Lenz, G. F. Study of the efficiency of natural polymer
extracted from the mandacaru cactus (Cereus jamacaru) as an aid in the coagulation
and flocculation processes in water treatment. Journal of Environmental Studies
(Online), v.14, n.2 esp, p.75-83, 2012.
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 4
ABSTRACT
*
Corresponding Author Email: [email protected].
82 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
1. INTRODUCTION
A comparison of source water quality and the desired finished water quality is
essential for any treatment process selection. However, flocculation is one of the most
common unit operations and is frequently employed for water and wastewater treatment.
Water treatment includes essentially three separated and sequential steps: coagulation
formation, particle destabilization, and interparticle interactions. It is an important step
for the removal of organic (viruses, bacteria, algae, protozoan cysts and oocysts,
microplastics, humic acids, particulate and dissolved organic matter that may be present
as natural organic matter - NOM) and inorganic particles (clay, silt, mineral oxides, and
erosion particles). The removal of particles is very important because they cause turbidity
(reduce water transparency and fotosintetic activities), may increase color to water;
promote infectious agents (microorganisms), and have toxic compounds adsorbed to their
external surfaces. NOM removal is quite important due the possibility of desinfection by-
products formation when chlorine is used for disinfection.
Ferric iron and aluminum salts are the most common salts used to promote the
coagulation and have been largely used in water and wastewater treatment plants.
Polymers may be or not associated in order to increase floc density and floc strength and
improve solid-liquid separation rates. Besides the addition of chemicals flocs formation
are extremely dependent on the hydrodynamic conditions. The understanding of all the
phenomena and mechanisms involved during flocculation is crucial to design efficient
flocculation units that may be applied to water or wastewater treatment. Thus, the main
goal of this chapter is to describe these mechanisms and applications of new flocculation
units for water and wastewater treatment.
The stability of the colloidal systems can be explained in part by the balance between
the London and van-der-Waals forces and the electric forces between the double layer of
particles (repulsion energy) known as DLVO theory, in honor of Derjaguin-Landau and
Verwey-Overbeek, a pair of Russian and Dutch scientists, respectively, who in the 1940s
independently developed this theory. However, from the 1980s onwards, with the
development of more advanced techniques (atomic force microscopy, for example), it
was possible to obtain results of the surface forces in aqueous medium, which
demonstrated (verified and proved the existing theoretical models) the existence of
Flocculation: Mechanisms and Applications for Wastewater Treatment 83
additional forces of hydration (repulsive force) and hydrophobic forces (attractive force
much greater than the van der der Waals forces), which were not predicted by the
classical theory. The inclusion of energy due to these forces, also known as structural
forces, resulted in a more modern concept named extended DLVO theory, or, simply, X-
DLVO (Yoon and Ravishankar, 1994; Israelachvili, 1992; Lins and Adamian, 2000;
Bratby, 2006; Bolto and Gregory, 2007; Vincent, 2012).
The particles generally have surface charge in aqueous media, which may originate
from the ionization of surface groups or sites, imperfection of the crystalline structure of
the solid surface, specific adsorption of ions and/or differentiated solubilization between
cations and anions. The surface potential of the colloids, the distribution of ions in
solution and the thermal effects lead to the formation of the double electric layer, shown
in Figure 1.
The double electric layer is modeled as being composed of two regions separated by
the Stern Layer (SL). The inner layer is known as the Stern Layer and the outer layer as
Gouy-Chapman or diffuse layer. In the presence of ions that are adsorbed specifically by
the chemical mechanism, the presence of two other planes is defined: the
Internal Helmholtz Layer (IHL), with potential I, and the External Helmholtz Layer
(EHL), with potential E. In the IHL the specific adsorption of ions by the chemical
mechanism occurs. The adsorption of co-ions, with charge of equal signal to
that of the surface of the particle, promotes an increase of the potential of the double
electric layer.
The adsorption of positive ions, more common case, promotes a decrease of the
electric potential or even the reversion of the charge of the particle. On the other hand, in
the EHL the ions are adsorbed by the physical or electrostatic mechanism, which, at the
most, promote neutralization of the electrokinetic potential of the colloid (). Because of
the difficulty of determining the surface electric potential of the particle, it is common
practice to measure the potential in the shear plane between the moving particle and the
surrounding liquid. The potential in this plane is known as Zeta Potential () or
electrokinetic potential.
3. PRINCIPLES OF FLOCCULATION
Stern Layer
Diffuse Layer
Ions in equilibrium
with bulk solution
different potential, being more complex than homocoagulation, especially when the
particles have negative surface charge. Thus, electrostatic interaction is the main
aggregative force acting on heterocoagulation.
When aggregation occurs with oils the process is known as agglomeration and,
because the agglomerates acquire spherical shape, it is also termed spherical
agglomeration. Through this mechanism hydrophobic particles present in water or
hydrophobized with residual surfactants may be agglomerated by the addition of a
nonpolar oil. This process consists of two stages:
86 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
(a) Oil-particle interaction: being the oil little soluble in water, the interaction occurs
between the droplets dispersed in water and the surface of the particles. If the
particles are hydrophobic, the oil spreads initially forming a lens and then a
liquid film that recovers its surface, making them more hydrophobic. Therefore,
the interaction is of hydrophobic character (hydrophobic forces) (Israelachvili,
1992);
(b) Capillary effect: when the oil concentration is high, the droplets occupy the entire
internal area available between the particles. At this stage, the capillary effect is
maximal and defines the spherical shape of the agglomerate with the particles
being held together by oily bridges. The formation of oil capillaries between the
particles promotes bonding and an increase in the hydrophobicity of the flocs.
The use of synthetic polymers in the solid-liquid separation rather than the coagulant
electrolytes allows a more effective process, providing more resistant aggregates (flocs),
higher sedimentation rates and more permeable filtration pans (Metcalf and Eddy, 2003;
Sincero et al., 2003; Fleer, 2010). The flocculating polymers employed for colloidal
destabilization include natural and synthetic products. Among the natural ones are
polyacrylamides, starches, proteins, tannins, biopolymers, guar gums and derivatives of
natural products, such as dextrin and sodium alginate (Metcalf and Eddy, 2003;
Schwoyer, 1981; Bratby, 1980). Most commercial polymers fits as synthetic polymers,
such as, for example, polyacrylamides and polyamides, or nonionic polymers such as
ethylene polyoxide (POE) and polyvinyl alcohol (PVA). As for the charge, the
flocculating polymers may be cationic (radical - NH3+), anionic (-COOH- radical),
nonionic (such as ethylene polyoxide), or amphoteric (semi-hydrolyzed polyacrylamides
which have negative and positive charges in the same jail). Most polymers are
hydrophilic, however, the presence of hydrophobic polymers (such as polyethylene oxide
and polyvinyl alcohol) may occur. The polymers may be low (10,000-100,000) and high
(> 100,000) molecular weight, reaching a molecular length of up to 1000 Å (Schwoyer,
1981; Bratby, 2006; Bolto and Gregory, 2007; Fleer, 2010).
The aggregation of the particles by polymeric bridges is called flocculation. The
polymer adsorb at the solid-liquid interface (hydrogen bonds, hydrophobic forces and
electrostatic attraction) by the electrostatic attraction mechanisms, polymer bridges or the
entrapment of the particles in polymer networks.
The kinetics of the formation of the flocs depends on the following steps:
one or more active sites, leaving the rest of the chain free, extended in the
solution;
(b) Conformation of the polymer forming loops, tails and trains. The conformation
of the adsorbed polymers depends on chain size (molecular weight), chain
flexibility, charge density (% hydrolysis), interaction energy between the
polymer and the colloid, the chemical and physical nature of the superficial sites
of the particles and competition between the polymer and other molecules
present in the solution;
(c) Adsorption of loops and tails and formation of polymer bridges. The
strength of the flocs depends on the number of bridges formed, and hence on the
number of loops and tails available. A factor of crucial importance
is the availability of sites in the particles to accommodate the bonds of
neighboring particles;
(d) Growth of the flocs under slow stirring. According to some authors (Arboleda,
1973; Bratby, 1980; Metcalf and Eddy, 2003; Bratby, 2006), after addition of the
destabilizing agent, rapid mixing is required for diffusion to take place in the
solid-liquid suspension and formation of the primary flocs. After the
appearance of the primary flocs, a slow mixing stage is usually required for
growth and formation of larger flocs and subsequent sedimentation.
However, flotation separation does not require the formation of large flocs. The
energy for the aggregation process is provided by the induction of velocity
gradients within the system (orthokinetic aggregation). The main
parameters involved in orthokinetic energy are the velocity gradient applied and
the stirring time.
4. PRINCIPLES OF FLOCCULATION
HYDRODYNAMIC MIXING
(a)
(d) (c)
𝑑𝑁
− 𝑑𝑡
= {𝐶𝑜𝑙𝑙𝑖𝑠𝑖𝑜𝑛 𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦}𝑥{𝐶𝑜𝑙𝑙𝑖𝑠𝑖𝑜𝑛 𝑓𝑟𝑒𝑞𝑢𝑒𝑛𝑐𝑦} = 𝛼 𝑥 𝐽 (1)
Microflocculation is the term used to denote particle aggregation due to the random
motion of molecules in the fluid. This random spatial motion of the molecules in the fluid
is also known as Brownian motion. The pericinetic aggregation starts immediately after
Flocculation: Mechanisms and Applications for Wastewater Treatment 89
destabilization and stabilizes within seconds, since it is significant for particles in the
range of 0.001 to 1 μm.
Macroflocculation is the term that refers to the aggregation of particles larger than 1
or 2 μm. Macroflocculation may occur due to the induced velocity gradient or differential
sedimentation. The energy imposed by a mixer is dissipated through velocity gradients
and the energy rate introduced is proportional to the set velocity gradient. The velocity
gradient is symbolized by G and is used to measure the mixing intensity. A high value of
G means an intense mixture, and a low value of G denotes a slow mixture. According to
Thomas et al. (1999), for a given value of G, orthokinetic aggregation is the predominant
mechanism when the particles exhibit uniformity of size, whereas the differential
sedimentation predominates when the particles present significant disparity of sizes. The
balance between the G employed and the mixing time can be expressed by the Camp
number (G.t).
In any unstable state, the composition of the liquid mass varies with time. This is the
case of most reactors used in treatment plants (mixers, flocculators, decanters, etc.) where
at any point it is found that both the velocity and the composition change constantly
because the water does not flow homogeneously, from the inlet to the outlet, i.e., not all
the flow entering the initial time reaches the outlet, exactly at the nominal holding time td
(Arboleda, 1973).
According to Bratby (1980); Bratby (2006) and Bolto and Gregory, 2007, the ideal
mixing type is the plug flow type, where all particles have the same residence time. In
mixers of the full mix type, some particles are short-circuited and others have very long
residence times. Such distribution of residence times is not desirable in the use of
hydrolyzable salts or polymers. In the case of the hydrolyzable salts, a short residence
time does not allow the complete adsorption of the hydrolyzed species on the surface of
the particles. The same occurs in the adsorption of polymers. On the other hand, intense
mixing for a very long time can break the polymer bridges between the particles and even
the polymer.
5. FLOCCULATION PROCESSES
Usually, three types of flocculators are employed in water and wastewater treatment:
hydraulic, mechanical, and pneumatic (Sincero et al., 2003). Hydraulic flocculators take
advantage of the energy that the flow acquires when flowing through a conduit to agitate
the liquid mass. Also, according to the direction of the liquid stream, inside the chambers,
hydraulic flocculators are divided into: horizontal flow, vertical flow and helical
flow.The mechanical flocculators are classified according to the type of movement of the
agitators, and can be alternating or rotating (Arboleda, 1973). Pneumatic flocculators
employ air to promote agitation. Table 1 shows the main classification of flocculators.
90 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
Hydraulic Flocculators
Hydraulic flocculators utilize the kinetic energy that the flow acquires when flowing
through a conduit to agitate the liquid mass. The most commonly hydraulic flocculator
models are the baffles (Arboleda, 1973; Sincero, et al. 2003).
Baffled Flocculators
The baffle flocculators consist of tanks, provided with internal channels, in which
water flows at a fixed velocity, producing some turbulence, at each change of flow
direction, making a 180° turn at the end of each channel, shown in Figure 3.
The most common are horizontal flow and vertical flow. A head loss occurs mainly
because:
(i) greater head loss (greater velocity gradient) in the 180° flow turns compared to
the straight stretches;
(ii) in the case of fixed baffles, the velocity is constant for each flow. In case of
variations in flow rate, the speed also changes, being very high or very low.
Outlet
Inlet
Alabama Flocculator
The Alabama Flocculator is made up of compartments interconnected by the bottom
through 90° curves facing upwards. The flow can be up and down inside the same
compartment. Removable nozzles installed at the exit of the reactor allow adjusting the
speed to the conditions of calculation or operation (Richter and Netto, 1991). According
to Vianna (1997), the flocs brought by the upstream tributary current collide with those
carried by the effluent stream, descending, resulting in the growth of the flocs. Thus,
according to the same author, it allows the existence of fewer chambers than vertical
baffle flocculators. Figure 5 illustrates an Alabama flocculator.
Cox Flocculator
This flocculator was developed by Cox for the Public Health Services Foundation of
Brazil (Richter and Netto, 1991). According to Vianna (1997), this type of flocculator has
a small number of chambers, and the interconnections between the chambers alternate
upper and lower positions, as shown in Figure 6. The main advantage of this type of
flocculator is the small number of compartments and as the main disadvantage, there is
cited the unevenness of the degree of agitation conferred on the liquid mass.
Table 2 summarizes the major advantages and disadvantages of hydraulic
flocculators.
Flocculation: Mechanisms and Applications for Wastewater Treatment 93
Inlet
Treatment Flux
Outlet
Advantages Disadvantages
Flow close to plug flow There is no flexibility in changing G
Does not require complementary Usually occupies a large space
devices
It does not require electrical energy
Ideal for small installations
Mechanical Flocculators
Vanes
These rotating flocculators have a system of vanes adhered by a horizontal or vertical
axis, which rotates by an electric motor, displacing the water and producing work. These
stirrers may have two, three or four arms, and, on each arm may have two or more vanes
or impellers joined by a central axis, as shown in Figure 7.
94 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
6. PNEUMATIC FLOCCULATORS OR
AIR-ASSISTED FLOCULATION
The necessary stirring for flocs formation can be accomplished by the introduction of
air into the system. The difference in density between the air bubbles and the water
causes the bubbles to rise to the surface. As the bubbles rise, they cause the water to
move, and consequently the mixture required for flocculation. In this case, the velocity
gradient reached can be controlled by adjusting the airflow. In published pneumatic
flocculation studies, chambers or columns are usually used where air is injected through a
diffuser (McConnachie, 1984; Sincero et al., 2003).
The in-line pneumatic flocculation of a biphasic flow (air-water), promoted by
mixers, can be explained by mechanisms similar to those of pneumatic flocculation in
chambers. In this process, the agitation necessary to promote flocculation is increased by
the injection of air through diffusers into the stream containing the effluent to be treated
and the destabilizing agent. In this way, the air released in the form of bubbles, in
addition to promoting a piston-type mixing in in-line mixers, has a high flocculation
efficiency for the removal of impurities that are suspended (Sholji and Kazi, 1997).
Through the principles of pneumatic flocculation, many authors have suggested a
new proposal for the generation of aerated flocs, in which dispersion of the air injected in
the form of small bubbles that adhere and/or trapped themselves to the flocs during its
formation (Owen et al., 1999; Fan et al., 2000; Da Rosa, 2002; Jameson, 1999; Zhao,
2002). In this way, aerated flocs “float” and are more easily separated later in a
96 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
𝑔.𝐷𝑏(𝜌𝑙 −𝜌)
𝐺𝑚𝑎𝑥 = 6.𝜇
(2)
where: ρ = density of the liquid (kg.m-3); ρ1 = density of the liquid; ρ = air in the bubble
(kg.m-3); Db = bubble diameter (m); g = acceleration of gravity (9.81 m /s2); μ = absolute
viscosity (10-3 N.s.m-2, for water at 20 ° C).
Table 4 shows the Gmax values for a bubble size range.
However, according to Masschelein (1992) cited by Metcalf and Eddy (2003) the
mean bubble diameter formed in this process is 5 mm, in a mean air flow of 10 percent of
the liquid flow. The velocity gradient due to the formation of bubbles in this range ranges
from 200 to 8200 s-1. In the case of extremely small bubbles of the order of micrometers,
the value of Gmax will also be small, becoming practically negligible. Table 5 presents
the main advantages and disadvantages of pneumatic flocculation.
Flocculation: Mechanisms and Applications for Wastewater Treatment 97
7. FLOCCULATION APPLICATIONS
Considerations about the main applications of the flocculation considering water and
wastewater treatment are discussed considering a compact helical unit named here as
Flocs Generator Reactor that may be applied for water and wastewater treatment
(emphasis is given for physicochemical processes) and a biological process application
for sewage treatment named here as bioflocculation for nitrogen uptake.
Curved configurations of circular tubes are widely used in heat exchangers, chemical
reactors, reverse osmosis units, blood oxygenation membranes, enzyme polymerization,
etc. These helical mixing units present enormous advantages over straight-line systems or
mechanical mixers, mainly due to hydrodynamic characteristics (higher Reynolds
numbers and higher velocity gradients) and secondary flow. This flow presents an action
of centripetal forces, with a movement along the walls and near the center of the tube,
increasing the resistance to the flow (Streeter, 1961; Berger et al., 1983, Agrawal and
Nigam, 2001; Elmaleh and Jabbouri, 1991; Ødegaard et al., 1992; Buchanan et al., 1998;
Carissimi and Rubio, 2005; Carissimi et al., 2007, Carissimi and Rubio, 2015; Carissimi
et al., 2018).
The evaluation of curved systems employed industrially and the analysis of the
different aggregation equipment used in the water and effluent treatment plants cited by
Carissimi and Rubio (2005) allowed the design and the design of a hydraulic aggregation
system in line with a piston regime, called Reactor Generator - RGF. Figure 9 shows the
response curve generated by the instant introduction of a tracer (methylene blue) for
determining the axial flow rate of the RGF. The pulse shown shows a slow tracer
spreading at the feed rate of 3 Lmin-1, and the time of 25 seconds denotes the tracer
residence time inside the RGF.
The FGR consists of an in-line helical mixing reactor for the solid-liquid aggregation
and separation of suspended particles, with a Brazilian Patent (INPI PI 0406106-3). The
necessary stirring for dispersion of the destabilizing agent and generation of the
aggregates is carried out by exploiting the kinetic energy of the hydraulic flow along the
helical tubular reactor.
98 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
Figure 9. Curve of the tracer response along the FGR. Conditions: feed flow =
3 Lmin-1, [AM] = 10000 mgL-1; FGR with a length of 12 m and a diameter of 2.5 cm.
Length
comprimento
microbolhas
Air Outer
diâmetro
de ar
microbubbles diameter
externo
Inlet
entrada
saída
Outlet
efluente + polímero
Water/wastewater + polymer
direção do fluxo
Flow direction
One (semi-pilot) unit of the FGR is shown in Figure 10; consisting of a transparent
polyurethane tube with internal diameter of 0.0125 m wrapped in the outside of a fixed
polyvinyl chloride (PVC) column, with an internal radius of 0,05 m, consisting of 32
rings, 12 m volume of 1.2 L, occupying a surface area of 0.60 mx 0.13 m. This model has
the alternative of injecting air microbubbles, generated through an air depressurising
system (such as in the dissolved air flotation process), making the reactor as an aerated
float floater (detailed later).
The use of online aggregation (pipelines, ducts, pipes) is not a common practice,
however, already existing in some industrial plants, either by the practical necessity of
operators, space optimization or by commercialization by companies specialized in water
and effluent treatment.
Flocculation: Mechanisms and Applications for Wastewater Treatment 99
Bioflocculation combines the primary influent of wastewater and the return activated
sludge from the secondary treatment in order to improve the nitrogen removal efficiency
as well as the nitrogen recovery.
Excess Nitrogen in rivers and streams causes water pollution and hence affects the
public health (Manuel, 2014). Nitrogen effluent discharge from wastewater treatment
plants are therefore becoming increasingly strict (Agency, 2018). The need for effective
nitrogen removal strategies to meet the nitrogen effluent limits while maintaining
minimal operating Wastewater Treatment Plants (WWTP) costs is therefore increasing
(Stare, Vrečko, Hvala, & Strmčnik, 2007). Conventional biological nitrogen removal
processes are usually conducted in a series of anoxic/aerobic reactor (Metcalf Eddy,
2014). The biological processes that remove nitrogen are nitrification and denitrification.
During the nitrification stage, high energy use for aeration is required to oxidize the
ammonia to nitrite then nitrate, making it a cost burden to WWTP. Nitrogen removal
requires 12 kWhel per person per year and significantly contributes to the overall
wastewater energy budget (Verstraete & Vlaeminck, 2011).
Nitrous oxide (N2O), which is a potent greenhouse gas, contributing for 5.3% of the
global anthropogenic greenhouse gas emissions in 2013 (EPA, 2014), can be emitted
during the nitrogen removal process in wastewater treatment. During the biological
nitrogen removal process, N2O is produced by nitrite reduction in the aerated
compartments (Toyoda et al., 2011) and during the anoxic stages as an intermediate
product (Colliver & Stephenson, 2000). The Environmental Protection Agency of the
United States reported that N2O from the wastewater sector accounts for about 3 per cent
of N2O emissions from all sources and ranks as the sixth largest contributor in the United
States (EPA, 2014). The increased presence of this gas in the atmosphere causes a rise in
the equilibrium temperature of the earth, thus arousing climate change. Therefore, the
need of an efficient nitrogen removal system that achieves low N2O emissions while
maintaining the nitrogen removal standards, is required.
The bioflocculation process is a proposed application, where the waste activated
sludge from the secondary treatment is used as flocculent. This process has proven to
enhance the nitrogen removal and recovery during the primary treatment stage. The
removal of nitrogen and organics from the primary effluent could reduce the energy
requirements in the activated sludge process and also reduce the N2O emissions produced
during the nitrogen removal processes. The biogas and the residual solids from the
facility’s anaerobic digester will increase upon digestion of primary sludge, which has
high energy and nitrogen content.
The bioflocculation application was tested using the influent wastewater and the
activated sludge obtained from the Narragansett Bay Commission Wastewater Treatment
Facility, located at Field’s Point in Providence, Rhode Island. The facility has an average
100 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
daily flow of 40 MGD providing services to about 226,000 people. This plant uses the
Integrated Fixed Film Activated Sludge (IFAS) process for the secondary treatment to
promote nitrification.
The influent wastewater used in this experiment was obtained from the outlet of the
primary treatment, while the activated sludge was obtained from the Return Activated
Sludge (RAS), which should have the same characteristics of the waste activated sludge.
The collected samples were tested and analyzed at the environmental engineering
laboratory located at the University of Rhode Island.
In order to apply the concept of bio flocculation process at this existing wastewater
facility with minimal investment costs, a pipe and a pump have to be installed to deliver
the secondary activated sludge back into the primary clarifier.
Figure 11 shows a schematic drawing of the proposed treatment system. The
recirculation of the activated sludge to the inlet primary treatment is shown in red.
The tests were conducted in sealed 1000 mL graduated cylinder. Influent wastewater
(or sewage) and activated sludge were mixed in the cylinders according to the test
conditions summarized in Table 6. The graduated cylinders were shaken to assure a good
mixing of the components, followed by 30 minutes of contact time between the two
materials. Bioflocculation efficiency was determined by the removal percentages of
nitrogen from the influent.
The removal percentage of organic matter takes into account the removal
mechanisms from both the influent and the activated sludge. This method was based on a
research study where settleable solids removal from bioflocculation was assessed
Flocculation: Mechanisms and Applications for Wastewater Treatment 101
(Araneda, Pavez, Luza, & Jeison, 2017). However, this method was modified for the
purpose of this application to calculate the total removal organic matter from the influent.
The results show that there is a considerable removal efficiency of respectively 59%,
47% and 25% for COD, TKN and NH3 concentrations in the second mixture (500:500),
where the sludge and sewage are mixed in the same ratio. The N2O emissions at the NBC
Field’s Point plant were found to be reduced by 1.4 µmol N2O/m2.s.
Anaerobic digesters are one of the most common waste sludge treatment. In this
process, the biosolids are broken down in the absence of oxygen, and partially converted
into gases, while the remainder is dried and becomes a residual soil-like material (Metcalf
Eddy, 2014). The treated sludge contains useful concentrations of nitrogen, phosphorus
and organic material.
When applying the suggested enhanced primary treatment using the bioflocculation,
the total nitrogen removed from the primary effluent will be transferred to the primary
sludge. The increase of the nitrogen availability in the primary sludge by 47% will
increase the benefits of the treated sludge, as it will have higher nitrogen contents, and
hence more beneficial to the grassland. The organic matter in sludge can also improve the
water retaining capacity and structure of some soils, especially when applied in the form
of dewatered sludge cake (Natural Resources Management and Environment
Department). The nitrogen being transferred to the primary sludge, is hence recovered to
a more useful way than being consumed by microorganisms in the secondary treatment,
which requires high amounts of energy too.
Anaerobic digestion reduces the volume of sludge and produces methane containing
biogas, a renewable energy, as a by-product. Collecting the energy from the biogas
through an efficient combustion process is both economically and environmentally
beneficial. Based on the results of the bioflocculation application, the percent of COD
concentrations in the influent that is transferred to the primary sludge was found to be
59%. The methane production achieved by primary sludge was found by (Mahdy,
Mendez, Ballesteros, & González-Fernández, 2015) to be 266 mL per gram of COD.
And, the methane has a calorific potential of 6.22E-3 kW.h/L CH4 (Metcalf Eddy, 2014).
By using these information, the amount of CH4 produced per day will increase by 76,021
cubic feet, which is around 51% increase of the usual methane production at the NBC
facility.
The bioflocculation application of the influent sewage can be advanced primary
treatment process. Its mechanism for the removal of organics, without any addition of
chemicals, enhances the nitrogen removal and recovery from wastewater treatment
plants. The experiments showed that TKN in the influent was reduced by 46% and the
NH3-N was reduced by 24% as a result of mixing the activated sludge from secondary
treatment with the primary influent, for a flocculation time of 30 min.
The removal performance of the bio flocculation in the primary treatment, has also
shown some energy consumption savings in the treatment aeration process. For the same
102 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
removal percentages mentioned above, the expected reduction in the aeration process was
about 18% for the cost and 36% for the air requirements.
Another main advantage of the bio flocculation process is the recovery of nitrogen
and organics in the primary sludge. The primary sludge, which is rich in nitrogen content,
can be used as fertilizer to supply agriculture soils instead of chemical fertilizers that
require large amounts of energy to produce. Also, the primary sludge which has a higher
energy content than the sludge removed at the secondary treatment due to the high
organics content, produce more biogas during digestion.
Finally, the calculations from the bio flocculation process have proven a reduction of
the nitrous oxide emissions from wastewater treatment due to the reduction of Nitrogen
content. In the enhanced primary treatment where the bio flocculation was applied, the
N2O emissions were reduced by 27%.
In conclusion, the Bio flocculation process has proven to be a new approach in
primary treatment of wastewater that leads to an energy efficient and improved process
design in wastewater treatment sector.
REFERENCES
Adamson, A. W. and Gast, A. P. (1997). Physical chemistry of surfaces. New York: John
Wiley and Sons, 6a. edição, 784 p.
Agency, U. S. E. P. (2018). The Sources and Solutions: Wastewater. Retrieved from
https://www.epa.gov/nutrientpollution/sources-and-solutions-wastewater.
Agrawal, S. and Nigam, K. D. P. (2001). Modelling of a coiled tubular chemical reactor.
Chemical Engineering Journal, v. 84, 437-444.
Araneda, M., Pavez, J., Luza, B. and Jeison, D. (2017). Use of activated sludge biomass
as an agent for advanced primary separation. Journal of Environmental Management,
192, 156-162. doi:10.1016/j.jenvman. 2017.01.030.
Arboleda, V. J. (1973). Teoría, diseño y control de los procesos de clarificación del água
[Theory, design and control of water clarification processes]. Lima, CEPIS, 558 p.
(In Spanish).
Berger, S. A., Talbot, L. and Yao, L. S. (1983). Flow in curved pipes. Annual Review of
Fluid Mechanics, 15, 461-512.
Bolto, B. and e Gregory, J. (2007). Organic polyelectrolytes in water treatment. Water
Research, 41(11), 2301-2324.
Bratby, J. (1980). Coagulation and flocculation. Uplands Press Ltd, England, 354 p.
Bratby, J. L. (2006). Coagulation and flocculation in water and wastewater treatment.
International Water Association, 450 pp.
Flocculation: Mechanisms and Applications for Wastewater Treatment 103
Buchanan, I. D., Nicell, J. A. and Wagner, M. (1998). Reactor models for horseradish
peroxidase-catalysed aromatic removal. Journal of Environmental Engineering,
794-802.
Carissimi, E., Miller, J. D. and Rubio, J. (2007). Characterization of the high kinetic
energy dissipation of the Flocs Generator Reactor (FGR). International Journal of
Mineral Processing, v. 85, 41-49.
Carissimi, E. and Rubio, J. (2005). The flocs generator reactor-FGR: a new basis for
flocculation and solid-liquid separation. International Journal of Mineral Processing,
75 (3-4), p. 237-247.
Carissimi, E. and Rubio, J. (2015). Polymer-bridging flocculation performance using
turbulent pipe flow. Minerals Engineering, 70, 20-25.
Carissimi, E., Sanagiotto, D. G., Camaño-Schettini, E. B. and Rubio, J. (2018). Revisiting
Coiled Flocculator Performance for Particle Aggregation. Water Environment
Research, v. 90, 322-328.
Colliver, B. B. and Stephenson, T. (2000). Production of nitrogen oxide and dinitrogen
oxide by autotrophic nitrifiers. Biotechnology Advances, 18(3), 219-232.
doi:10.1016/S0734-9750(00)00035-5.
Dobiás, B. and e Stechemesser, H. (2005). Coagulation and flocculation. Marcel Dekker
Inc, New York, 882 pp.
Elmaleh, J. and Jabbouri, A. (1991). Flocculation energy requirement. Water Research,
25 (8), p. 939-943.
EPA, U. S. E. P. A. (2014). Inventory of U.S. Greenhouse Gas Emissions and Sinks:
1990-2012. The Air Pollution Consultant, 24 (3), 1_17.
Fleer, G. J. (2010). Polymers at interfaces and in colloidal dispersions. Advances in
Colloid and Interface Science, 159(2), 99-116.
Israelachvili, J. N. (1992). Intermolecular and surface forces. Academic Press, University
of California, 2nd Edition, 450 p.
Lins, F. F. and Adamian, R. (2000). Minerais coloidais, teoria DLVO estendida e forças
estruturais [Colloidal minerals, extended DLVO theory and structural forces]. Rio de
Janeiro, CETEM/MCT, 29 p. (In Portuguese).
Mahdy, A., Mendez, L., Ballesteros, M. and González-Fernández, C. (2015). Algaculture
integration in conventional wastewater treatment plants: Anaerobic digestion
comparison of primary and secondary sludge with microalgae biomass. Bioresource
Technology, 184, 236-244. doi:10.1016/j.biortech.2014.09.145.
Manuel, J. (2014). Nutrient pollution: a persistent threat to waterways. Environmental
health perspectives, 122(11), A304. doi:10.1289/ehp. 122-A304.
Metcalf, e Eddy. (2003). Wastewater engineering: treatment and reuse. Editores:
Tchobanoglous, G.; Burton, F. L.; Stensel, H. D. Metcalf e Eddy, Inc., McGraw Hill,
4th Edition, 1819 p.
104 Elvis Carissimi, Cristiane Oliveira Rodrigues, Dounia Elkhatib et al.
Chapter 5
FLOCCULATION OF AOM
IN WATER TREATMENT
ABSTRACT
*
Corresponding Author Email: [email protected].
108 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
common impurities, both of organic and inorganic nature. Mutual interactions have been
proven, while their influence on flocculation efficiency can be either positive or negative,
depending on the AOM character, pH conditions and on the ratio between AOM, the
other polluting agents and coagulants. Finally, AOM also appeared to alter the properties
of flocs, with an impact on the subsequent separation steps. In further research, a
particular emphasis should be put on AOM components that are difficult to coagulate, the
interactions of AOM with other impurities and on elucidation of the relationship between
AOM and floc properties.
1. INTRODUCTION
1
The terms ‘flocculation’ and ‘coagulation’ commonly appear in the field of water treatment. These processes
result in clumping of particles/molecules and subsequent formation of flocs, while ‘coagulation’ rather stands
for the destabilization of particles (involving charge neutralization) and ‘flocculation’ for the aggregate (floc)
formation. However, the terminology is not consistent among different studies and the terms are being used in
Flocculation of AOM in Water Treatment 109
competing with target pollutants regarding adsorption onto activated carbon (Pivokonsky
et al., 2016). Flocculation followed by separation of the formed aggregates
(sedimentation, flotation, filtration) is a fundamental step in water treatment. It is
extensively employed at many drinking water treatment plants worldwide and is
irreplaceable due to its operational benefits and cost-effectiveness. It also improves
performance of potential downstream treatment processes, such as membrane filtration or
adsorption onto activated carbon (Zhang et al., 2014). Flocculation has a potential to
remove different polluting agents. Good removal efficiencies (94-99%) of algal and
cyanobacterial cells were obtained under optimized flocculation conditions (Henderson et
al., 2010; Baresova et al., 2017). However, AOM is generally much more difficult to
coagulate, and additionally, AOM also influences flocculation of other compounds
(Bernhardt et al., 1985; Henderson et al., 2010; Safarikova et al., 2013; Pivokonsky et al.,
2015, 2016; Baresova et al., 2017).
2. WHAT IS AOM
similar meaning. Thus, the terms ‘flocculation’ and ‘coagulation’ are considered equal in this chapter and the
term ‘flocculation’ is preferred.
2
As AOM is a mixture of various organic compounds, its concentrations are commonly being expressed as DOC
(dissolved organic carbon) in mg L-1 DOC.
3
Also the term ‘IOM – intracellular organic matter’ is sometimes being used in this sense (Pivokonsky et al., 2006;
Fang et al., 2010; Li et al., 2012). Additionally, ‘SOM – surface-retained organic matter’ is distinguished in
some studies (Takaara et al., 2010).
110 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
alter overall water quality (Zhang et al., 2010). Additionally, COM can also occur as a
result of water treatment processes that induce cell rupture, e.g., pre-oxidation (Ma et al.,
2012b; Coral et al., 2013).
4
recognized as Dolichospermum flos-aquae since 2009
Flocculation of AOM in Water Treatment 111
Due to its varying compositions, algal organic matter differs in several characteristics
that alter its flocculation behaviour. Apart from the division into peptides/proteins and
non-proteinaceous matter, properties of AOM that are substantially important from the
perspective of water treatment include (1) the degree of hydrophilicity/hydrophobicity,
(2) molecular weight (MW) distribution and (3) surface charge (Pivokonsky et al., 2016).
2.3.1. Hydrophilicity/Hydrophobicity
Algal organic matter is usually divided into three groups according to the level of its
hydrophobicity, i.e., into hydrophilic (HPI), hydrophobic (HPO) and transphilic (TPI)
fraction (Her et al., 2004; Henderson et al., 2008a; Pivokonsky et al., 2014; Baresova et
112 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Species EOMexp proportion (%) EOMstat proportion (%) COM proportion (%)
HPI HPO TPI HPI HPO TPI HPI HPO TPI
A. flos-aquaea - - - 81 8 - - - -
Aph. flos-aquaea - - - 63 18 - - - -
A. formosab 73 14 13 68 19 13 - - -
C. geitleric 71 21 8 73 22 5 89 10 1
C. vulgarisb 63 24 13 72 11 17 - - -
F. crotonensisc 74 17 9 74 19 7 90 8 2
Melosira sp.b - - - 60 32 8 - - -
M. tenuissimad - - - - - - 77 7 16
M. aeruginosac 69 27 4 69 28 3 87 12 1
S. subspicatusa - - - 54 26 - - - -
a
Goslan et al., 2017; bHenderson et al., 2008a; cPivokonsky et al., 2014; dBaresova et al., 2017.
The HPI fraction of organic matter include carbohydrates, amino acids and other
organic acids, aldehydes, ketones, etc. (Edzwald, 1993). HPO compounds (determined to
form approximately 8-32% of EOM and 7-12% of COM) include humic and fulvic acids,
aromatic acids and amines, and high-MW alkyl carboxylic acids (Edzwald, 1993;
5
SUVA (specific UV absorbance) corresponds to the aromaticity of a sample. It is defined as the UV absorbance at
a given wavelength (254 nm) normalized for DOC concentration.
Flocculation of AOM in Water Treatment 113
Pivokonsky et al., 2014; Baresova et al., 2017; Goslan et al., 2017). AOM
peptides/proteins contain both HPI and HPO parts because they are composed of
hydrophilic as well as hydrophobic amino acids. However, in aqueous solutions, the HPO
segments tend to cluster together, leading to protein folding, while the HPI regions are
exposed to the surroundings (Creighton, 1993).
Figure 1. Molecular weight fractionation of (a) COM peptides/proteins and (b) non-proteinaceous
cellular organic matter of different phytoplankton species (Chlamydomonas geitleri, Chlorella vulgaris,
Fragilaria crotonensis, Merismopedia tenuissima and Microcystis aeruginosa). Adapted from
Pivokonsky et al. (2014), Baresova et al. (2017) and Naceradska et al. (2018).
114 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
By contrast, substances with very low MW can also form a substantial part of AOM
(Hoyer et al., 1985; Henderson et al., 2008a). E.g., 30% of Chlorella vulgaris AOM, 38%
of M. aeruginosa AOM, 53% of Melosira sp. AOM and even 81% of Asterionella
formosa AOM was determined to be < 1 kDa (Henderson et al., 2008a). More
information on the proportion of high- and low-MW organics in AOM of different
species is shown in Figure 1, where the difference between AOM peptides/proteins and
non-proteinaceous matter is emphasized.
Figure 2. Molecular weight distributions of EOM and COM peptides/proteins of (a) Chlamydomonas
geitleri, (b) Fragilaria crotonensis, (c) Merismopedia tenuissima and (d) Microcystis aeruginosa. EOM
MWs were determined at exponential and stationary growth phases (EOMexp, EOMstat, resp.), COM was
obtained at the stationary growth phase. Adapted from Pivokonsky et al. (2014) and Baresova et al.
(2017).
The MW distribution differs not only between the diverse phytoplankton species, but
also depends on the age of the culture and AOM fraction (i.e., EOM vs. COM). When the
MWs of AOM peptides/proteins were analysed, COM was found to contain a much
Flocculation of AOM in Water Treatment 115
2.3.3. Charge
The charge of AOM is strongly pH dependent and is attributable to the presence of
diverse functional groups that can release or accept proton, depending on the pH
conditions (Creighton, 1993). It was found that AOM is negatively charged throughout a
wide range of pH values (Bernhardt et al., 1985; Henderson et al., 2008a), e.g.,
Henderson et al. (2008a) determined negative zeta potential values of EOM derived from
various species in the pH range 2-10. Nevertheless, the charge density varies among the
species and their growth stage (Bernhardt et al., 1985; Paralkar and Edzwald, 1996;
Henderson et al., 2008a). For example, the charge density of M. aeruginosa EOM at pH 7
decreased from 0.2 meq per 1 g DOC at the exponential growth phase to 0.1 meq per 1 g
DOC at the stationary growth phase. On the contrary, the charge density of C. vulgaris at
pH 7 increased from 0.9 to 3.2 meq per 1 g DOC from exp to stat phases (Henderson et
al., 2008a).
Additionally, distinct AOM constituents differ in their charge properties. AOM
peptides/proteins are amphoteric due to the content of diverse functional groups (both
acidic and basic, i.e., -OH, -COOH, -SH, -NH3+, =NH2+) (Creighton, 1993).
Peptides/proteins with isoelectric points (pI) ranging from 4.8-8.1 were identified in
COM of M. aeruginosa (Pivokonsky et al., 2012; Safarikova et al., 2013). Nevertheless,
negative charge of peptides/proteins prevails within pH ranges relevant for water
treatment (Pivokonsky et al., 2012). The charge of polysaccharides stems from the
presence of uronic acids, containing weakly acidic -COOH groups (Hoyer et al., 1985).
However, the non-proteinaceous fraction is generally considered to bear less ionisable
groups and thus is less charged. For example, COM peptides/proteins produced by M.
aeruginosa were reported to contain 100 mmol titratable functional groups per 1 g DOC
(Safarikova et al., 2013), while non-proteinaceous COM of Chlorella vulgaris possessed
about 14 mmol titratable functional groups per 1 g DOC (Naceradska et al., 2018).
Similarly, alginate, an anionic polysaccharide, appeared to involve approx. 10 mmol -
COOH groups per 1 g DOC (Gregor et al., 1996).
116 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
3. FLOCCULATION OF AOM
A variety of chemicals have been trialled for coagulating algae and AOM including
traditional metal coagulants aluminium sulphate (alum), ferric sulphate (FS),
ferric chloride (FC) and inorganic polymers, for example polyaluminium chloride (PACl)
and polyferric sulphate (PFS). Natural polymers, such as chitosan or cationic
starch, have also been tested for coagulation of cell + AOM mixtures (Cheng et al., 2011;
Vandamme et al., 2012, 2014; Garzon-Sanabria et al., 2013). In addition,
alkaline flocculation, i.e., by chemical co-precipitation with calcium and magnesium salts
of the culture medium under alkaline pH values, is used for algal biomass harvesting for
biofuel production (Vandamme et al., 2012; González-Fernándes and Ballesteros, 2013;
Vandamme et al., 2016; Branyikova et al., 2018).
Most of the investigations into the flocculation of algae-laden waters have focused on
algal and cyanobacterial cells (Henderson et al., 2008b, 2010; Cheng et al., 2011;
Vandamme et al., 2012; Wyatt et al., 2012; Garzon-Sanabria et al., 2013; Gonzalez-
Torres et al., 2014; Baresova et al., 2017). They found that algal cells easily combine
with commonly used Al-based or Fe-based coagulants through electrostatic bridging
between negatively charged algal cell surface and positively charged coagulant hydroxide
precipitates at about neutral pH values (Henderson et al., 2008b; Wyatt et al., 2012;
Gonzalez-Torres et al., 2014; Baresova et al., 2017; see Figure 3). At higher coagulant
doses, sweep flocculation has also been reported to be influential (Wyatt et al., 2012;
Gonzalez-Torres et al., 2014).
The negative surface charge of algal cells arises from deprotonation of acidic groups
(mostly -COOH) present at the cell surface or in EOM attached at the cell surface
(Henderson et al., 2008b; Wyatt et al., 2012; González-Fernándes and Ballesteros, 2013).
Wyatt et al. (2012) observed efficient flocculation of Chlorella zofingiensis cells even at
alkaline pH values, which may be attributed to interactions between cationic amine
Flocculation of AOM in Water Treatment 117
groups on the cell surface with anionic coagulant precipitates, or when using sweep
coagulation at higher coagulant doses (Duan and Gregory, 2003) or co-precipitation with
calcium and magnesium salts. When natural polymers are used as coagulants, different
coagulation mechanisms may be expected, as shown by Cheng et al. (2011) for
flocculation of cells of Chlorella variabilis by chitosan.
In this study, flocculation of Chlorella cells improved at pH 8.5 compared to pH 5.5
and 7, suggesting that hydrogen bonds between chitosan (a poly-glucosamine polymer
with an isoelectric point around 6.5) and cell wall polysaccharides may be more
important than electrostatic interactions. Several studies (Bernhardt and Clasen, 1991;
Henderson et al., 2008b, 2010; Wyatt et al., 2012; Zhang et al., 2012) revealed a strong
stoichiometric relationship between cell surface area and coagulant demand
for spherical cells. The presence of EOM usually leads to the higher coagulant demand
(Henderson et al., 2010; Ma et al., 2012b; Zhang et al., 2012), which is further elaborated
in section 4.3.
Compared to algal cells, flocculation of AOM is markedly less investigated and
provides substantially lower removal efficiencies (Bernhardt et al., 1985, 1986, 1991;
Widrig et al., 1996; Pivokonsky et al., 2009a, b; Henderson et al., 2010; Safarikova et al.,
2013; Pivokonsky et al., 2012, 2015, 2016; Baresova et al., 2017; Naceradska et al.,
2017; Tang et al., 2017; Naceradska et al., 2018).
Henderson et al. (2010), for example, observed good cell removal (94-99%) for
Microcystis aeruginosa, Chlorella vulgaris, Asterionella formosa and Melosira
sp. by alum, while their EOM removal efficiency was 46-71%. Tang et al.
(2017) recorded high removal efficiencies (> 90%) for cells of cyanobacterium
Microcystis aeruginosa coagulating with PACl, but low removals for dissolved EOM (<
10%) and for cell-bound EOM (< 40%). Similarly, Baresova et al. (2017) obtained 99%
cell removal efficiencies for cyanobacterium Merismopedia tenuissima and
ferric sulphate, while its COM was reduced by 43-53% depending on the initial COM
concentration.
Similar to cell removal, AOM surface charge that is closely related to pH value is
crucial for flocculation. AOM is usually removed at acidic pH values, at which AOM
negatively charged functional groups interact with positively charged coagulant
hydroxopolymers (Widrig et al., 1996; Hu et al., 2006; Pivokonsky et al., 2009a, b). To
illustrate, Widrig et al. (1996), who examined the removal of EOM derived from green
algae Scenedesmus quadricauda and Dictyosphaerium pulchellum and cyanobacterium
Microcystis aeruginosa by alum and ferric chloride at two pH values, 5 and 8, showed
that overall removals were improved at pH 5 (M. aeruginosa – 20%, S. quadricauda –
25%, and D. pulchellum – 50%) compared to pH 8 (5-10%). Likewise, Baresova et al.
(2017) obtained the highest coagulation efficiencies for M. tenuissima COM at pH
ranging between 5-6.5, while cells were preferably removed at pH 6-7.7, suggesting that
different coagulation mechanisms were employed. Cells probably interacted through
118 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
for COM peptides/proteins (Pivokonsky et al., 2012). The low removals were attributed
to high content of low-MW organics (< 10 kDa) in non-proteinaceous COM, which were
disinclined to flocculate.
Figure 4. Flocculation mechanisms of COM proteins and coagulant: (a) pH = 4-6, (b) pH = 6-8, low
COM/Fe ratio and (c) pH = 6-8, high COM/Fe ratio. Adapted from Pivokonsky et al. (2012).
The reluctance of low-MW organics to flocculate was also ascertained for COM
peptides/proteins (< 10 kDa) (Pivokonsky et al., 2012, 2015) and low-MW humic
substances as a part of natural organic matter commonly occurring in surface water
sources (Sillanpää et al., 2017). It should be noted that most cyanobacterial toxins are of
low-MW, for instance microcystins have MW around 1 kDa, and therefore coagulation is
largely ineffective for their removal (de Figueiredo et al., 2004).
When flocculation performance of the most used coagulants is compared, it can be
concluded that aluminium-based and ferric-based coagulants perform very similarly in
AOM coagulation in terms of removal efficiencies under slightly shifted optimum pH
values, given by different distribution diagrams of their hydrolysis products (Widrig et
al., 1996; Pivokonsky et al., 2009a; Safarikova et al., 2013).
Studies that dealt with the impact of AOM on coagulation of inorganic particles
showed that the AOM interacted with the inorganic particles and that the presence of
AOM substantially changed the optimum pH for effective flocculation by ferric or
aluminium coagulants (Bernhardt et al., 1985, 1986, 1991; Safarikova et al., 2013).
Inorganic particles are often present in raw water, where they cause high turbidity, and
they usually bear a negative charge in a wide range of pH values. The interactions
between clay surfaces and organics were demonstrated in soils for lots of organic
compounds, such as mono- and polysaccharides, amino acids, proteins, nucleic acids and
humic substances (Goring and Bartholomew, 1952; Greenland, 1956; Pinck, 1962; Parfit
and Greenland, 1970; Labille et al., 2005). Organic compounds that bear a charge (e.g.,
acidic polysaccharides, proteins) can interact with the negative charge of clays by
electrostatic interactions. These are affected by pH of the solution, which governs
dissociation constants of polar functional groups and thus the amount of charged groups
in both clays and organic compounds (Pinck, 1962). For neutral compounds, several
molecular mechanisms for adsorption via weak bonds were proposed: (1) substitution of
interlayer water molecules solvating exchangeable cations (Parfit and Greenland, 1970),
(2) hydrogen bonds between macromolecule hydroxyl and clay surface oxygens, or (3)
hydrogen bonds between hydroxyl groups on sheet edges and macromolecule (Labille et
al., 2005). Clay-organic interactions were ascertained also in cases of water treatment.
For instance, Bernhardt and co-workers (Bernhardt et al., 1985, 1986, 1991), who
investigated the effect of EOM of MW > 2 kDa excreted by several algal species on the
Flocculation of AOM in Water Treatment 121
coagulation of inorganic quartz particles, concluded that EOM polymers (neutral and
acidic polysaccharides) were able to attach to the negatively charged surface of quartz
particles by means of hydrogen and covalent bonds. Likewise, a study by Safarikova et
al. (2013) suggested that cellular peptides/proteins of M. aeruginosa interact
electrostatically (though -NH3+ and =NH2+ groups) with the negatively charged surface of
kaolin particles. In both cases, AOM and inorganic particles formed negatively charged
clusters, which were then removable by both ferric and aluminium coagulants in the pH
range, where Fe/Al formed hydrolytic products bearing a positive charge, i.e., at pH
ranges of 4-6.5 in the case of Fe and 4.5-7 in the case of Al (Figure 6).
Figure 6. Flocculation mechanism of kaolin, COM proteins and coagulant. Adapted from Safarikova et
al. (2013).
by EOM polymers enhances the stability of the dispersed system by steric and charge
stabilization (Bache and Gregory, 2007) and impairs flocculation. Similar results were
obtained also in the case of alginic acid, an polyuronic acid composed of two uronic acids
(poly-D-mannuronic acid and L-guluronic acid), which was used as a model compound
for EOM acidic polysaccharides (Bernhardt et al., 1985, 1986, 1991). Similar to the
flocculation of AOM solely (section 3.), Safarikova et al. (2013) ascertained that high-
MW peptides/proteins were removed by flocculation, while low-MW ones were reluctant
to flocculate.
In many regions, humic substances (HS), comprising humic and fulvic acids (HA and
FA), constitute the majority of natural organic matter in surface waters and may also
occur together with AOM (Knauer and Buffle, 2001). While a lot of emphasis has been
put on removal of HS, very little attention has been paid to possible HS-AOM
interactions and their simultaneous removal by flocculation. One of the rare
investigations was done by Jiang et al. (1993) who flocculated cells of diatom
Asterionella formosa and AOM excreted during its growth by four different inorganic
coagulants (polyferric sulphate, ferric sulphate, aluminium sulphate, polyaluminium
chloride) in the presence of HS. They ascertained that the addition of HS to the algae +
AOM solutions lead to a reduction of flocculation performance with increasing
concentration of HS. With a greater HS concentration, higher doses of coagulants were
required to achieve overall charge neutralization and removals comparable to those in the
absence of HS. On the other hand, several studies have demonstrated that HS is able to
adsorb onto the surfaces of freshwater phytoplankton cells and that the HS-cell
interactions are strongly dependent on pH value (Campbell et al., 1997; Vigneault et al.,
2000; Knauer and Buffle, 2001). Campbell et al. (1997) have proposed two mechanisms
for the adsorption of HS onto the phytoplankton cell surfaces: (1) hydrogen bonds and (2)
the formation of hydrophobic bonds between the cell surface and the hydrophobic
domain of the HS. Interactions between HS and microbial products, such as
polysaccharides and proteins, were also observed during membrane filtration (Myat et al.,
2014). Myat et al. (2014) investigated the impact of interactions between model
compounds, sodium alginate and bovine serum albumin (BSA) as representative
biopolymers, with humic acid on membrane fouling. They detected formation of
aggregates between sodium alginate and BSA with humic acid by liquid chromatography
and they also employed molecular dynamics simulations to provide insights into the
interaction mechanisms.
For the BSA-humic acid system, they showed that electrostatic, hydrophobic and
hydrogen bonding were the dominant types of interactions predicted. For the alginate-
Flocculation of AOM in Water Treatment 123
humic acid system, the interactions predicted were divalent ion-mediated interactions
only. The existence of interactions between HS and BSA was confirmed by Pivokonsky
et al. (2015). Moreover, they found that HS interact also with cellular proteins of M.
aeruginosa (MA) and that interactions between HS-BSA and HS-MA proteins have very
similar consequences for flocculation. In both cases, the clusters of HS-BSA and HS-MA
proteins were flocculated by alum with DOC removals of 83% and 80%, respectively, in
the pH range 5.5-6.2 through a charge neutralization mechanism, i.e., positively charged
Al-hydroxopolymers interacted with negatively charged functional groups of BSA/MA
proteins and HS (Figure 7).
Figure 7. Flocculation mechanism of humic substances, COM proteins and coagulant. Adapted from
Pivokonsky et al. (2015).
Most studies that dealt with the simultaneous flocculation of AOM and algal cells,
have focused on the flocculation of algal cells in the presence of extracellular organic
matter, either produced by cells during their growth into cultivation media in water
treatment or during the harvest of algal biomass (Bernhardt and Clasen, 1991; Henderson
et al., 2010; Vandamme et al., 2012; Garzon-Sanabria et al., 2013; Wu et al., 2012;
124 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
cells, COM served as a cationic polymer coagulation aid and that the presence of COM
shifted the optimum pH values for flocculation. While the cells were removed by
adsorption onto coagulant precipitates (Fe-oxide-hydroxides) at around neutral pH, the
COM + cell mixtures underwent charge neutralisation by Fe-hydroxopolymers within
moderately acidic pH (5-6.5). Furthermore, coagulant doses for COM + cell mixtures
were even slightly lower than those for single COM flocculation and the achieved DOC
removals for COM + cell mixtures (37-57%) were comparable to the single COM
flocculation (43-53%, rising with the increasing initial COM concentration). The high
cell removals (up to 99%) were obtained in both the presence and absence of COM. The
beneficial effect of COM on flocculation of M. tenuissima cells stemmed from bridging
cells by high-MW COM, mainly on basis of hydrogen bonds and electrostatic
interactions (Figure 8).
Figure 8. Flocculation mechanism of cells, COM and coagulant. Adapted from Baresova et al. (2017).
5. AOM AS A COAGULANT
Natural polymeric materials are increasingly used in water and wastewater treatment
as a coagulant or coagulant aid. The primary flocculation mechanism is ‘bridging’ which
involves adsorption of the polyelectrolyte molecule from solution onto the surface of
suspended particles. The most commonly used natural polymers are chitosan, alginate,
starch and extract from seeds of Moringa oleifera. Since alginates bear a strong
resemblance to algal products, it can be assumed that some compounds contained in
AOM may also serve as coagulants. For instance, Devrimci et al. (2012)
showed that calcium alginate proved to be a successful coagulant for turbid waters
containing clay (predominantly smectite) and that calcium ions eased alginate-clay
interaction.
126 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Figure 9. Flocculation mechanisms of (a) kaolin and COM proteins (adapted from Safarikova et al.,
2013), (b) humic substances and COM (adapted from Pivokonsky et al., 2015) and (c) cells and COM
proteins (adapted from Baresova et al., 2017).
Floc properties, such as size, structure, shape, strength or settling velocity, are
important parameters for the downstream treatment processes, i.e., separation by
sedimentation, flotation and filtration. Each of these technological steps requires different
flocs for their successful separation. Small (generally < 50-60 µm) and dense flocs are
suitable for a single-stage separation by sand filtration (Ngo et al., 1995; Bubakova and
Pivokonsky, 2012). For DAF (dissolved air flotation), Edzwald (1995) recommended
producing strong flocs with particle size distributions in the range 10-30 µm (or 25-50
µm in Edzwald, 2010); however, Vlaški et al. (1997) reported that larger flocs (> 50 µm)
formed by very low shear rate (G = 10 s-1) and hence having low density, resulted in
efficient DAF as well. Finally, sedimentation is generally considered to be suitable for
removal of large (> 100 µm) and dense flocs which have high settling velocities (e.g.,
Edzwald, 1995).
Generally, floc properties are influenced by several basic parameters: the
composition (type) and concentration of floc components, i.e., impurities to be removed
and the coagulant; the shear rate (velocity gradient, G) and the time of its action
(flocculation time). The properties of algal/cellular organic matter flocs have only been
examined marginally thus far; however, some works concerning AOM, sometimes
together with algal/cyanobacterial cells, have been published in recent years (Henderson
et al., 2006; Pivokonsky et al., 2009a, b; Gonzalez-Torres et al., 2014, 2017; Vandamme
et al., 2014; Jiao et al., 2015; Chekli et al., 2017; Filipenska et al., 2018).
Some of them examined either the influence of coagulant dose and type, or pH value
on AOM floc properties. To illustrate, Chekli et al. (2017) evaluated the performance of
TiCl4 and polytitanium tetrachloride (PTC) for the removal of AOM from a marine
diatom Chaetoceros muelleri in comparison with conventional FeCl3 treatment. They
found that PTC and TiCl4 reached higher removal efficiencies (in terms of turbidity,
DOC and UV254) and made larger, stronger and more compact flocs than FeCl3.
Gonzalez-Torres et al. (2014) examined the properties of flocs made of M. aeruginosa
cells and its AOM using Al2(SO4)3 or FeCl3 as coagulants at pH 6 (lower coagulant dose,
charge neutralisation mechanism) and 7 (higher coagulant dose, sweep flocculation
mechanism). Ferric flocs were confirmed to be larger than alum flocs regardless of the
pH value or coagulant dose. Within one coagulant type, the coagulant dose was another
factor influencing the floc size, since a higher dose produced larger flocs. In the case of
the constant both coagulant type and dose, the pH values also affected the size of flocs.
Specifically, pH 7 resulted in larger flocs than pH 6. Floc strength was determined
primarily by the coagulant dose, irrespective of coagulant type. Higher coagulant doses
produced stronger flocs, the strongest of which being produced at the higher dose of Fe at
pH 6.
Flocculation of AOM in Water Treatment 129
The others evaluated the influence of AOM on the properties of flocs made of other
impurities, such as clay minerals or cells. Unfortunately, a direct comparison of floc
properties reported by different researchers is made difficult because of the varying
experimental conditions (different/unknown shear rates, different coagulant doses or
various AOM compositions and concentrations). Yet some general conclusions can be
made. The principal finding is that the presence of AOM increases the floc size
(Vandamme et al., 2014; Filipenska et al., 2018). For instance, Vandamme et al. (2014)
coagulated Chlorella vulgaris cells with and without AOM using alum, chitosan and
cationic starch as coagulants, electro-coagulation and alkaline flocculation. Generally,
AOM increased the floc size (up to ten times for chitosan). However, flocs formed using
cationic starch remained of the same size irrespective of the AOM presence. A clear trend
was observed by Filipenska et al. (2018) who coagulated kaolinite and COM proteins of
M. aeruginosa (separately and together) using Al and Fe salts as coagulants. In all cases
COM proteins definitely caused the floc size to increase. However, the structure of flocs
was more influenced by the coagulant type than by the presence of COM as the most
compact and regular (in shape) flocs were Al-kaolinite and Al-COM flocs. Henderson et
al. (2006) reported that flocs created from C. vulgaris cells and EOM were larger, but
weaker, than flocs formed by inorganic particles (kaolin) or NOM (humic matter).
Unfortunately, the tests with single cells without EOM have not been made, therefore it is
not possible to evaluate the specific EOM contribution to the resulting floc size.
Moreover, Pivokonsky et al. (2009b), who coagulated COM of M. aeruginosa with ferric
sulphate, reported that COM concentration considerably affected the floc size, with lower
COM concentration producing smaller flocs. The influence of the origin and composition
of AOM on the floc characteristics can be demonstrated by research conducted by
Gonzalez-Torres et al. (2017). They coagulated cells and associated EOM of C. vulgaris
and M. aeruginosa by alum and used reflectance Fourier transform infrared (FTIR)
imaging for biochemical characterisation of the associated flocs. M. aeruginosa flocs
were measurably more uniform and smaller than C. vulgaris flocs and visibly more
dense, which was attributed to the difference in AOM concentration and composition.
The large flocs of C. vulgaris were characterized by a homogenous distribution of
proteins and polysaccharides across the floc, indicative of high glycoprotein content. In
contrast, the smaller but stronger flocs of M. aeruginosa had localized areas of increased
protein concentration at the edge of regions which were absent of proteins and
polysaccharides (and thus potentially comprising coagulant).
Flocculation mechanisms and inter-particle interactions (mentioned in
previous sections) acting within flocs also seem to influence the floc properties.
Filipenska et al. (2018) pointed out an interesting phenomenon. When floc size
130 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
and shear rate were plotted in a log-log plot6, a change in γ constant at approximately G ~
120 s-1 for all types of flocs and an additional change in γ for COM flocs,
between G of 40 to 50 s-1 and 60 to 70 s-1 for Al and Fe, respectively, could be observed
(Figure 10).
Figure 10. Floc size vs. shear rate log-log plots of six different types of flocs. A change in γ constant at
approximately G ~ 120 s-1 for all types of flocs and an additional change in γ for COM flocs, between G
of 40 to 50 s-1 and 60 to 70 s-1 for Al and Fe, respectively, can be observed. Adapted from Filipenska et
al. (2018).
CONCLUSION
This review shows that flocculation is a useful technology for the treatment of algae-
laden waters. Specific outcomes and further research needs can be summarized as
follows:
6
The dependence of the floc size on the shear rate is described by a decreasing power function (e.g., Parker et al.,
1972) dav/max = CG-2γ, where dav/max is the average or maximal diameter of flocs in the system, G is the shear
rate (global, average, mean) and C and γ are constants that include parameters such as the composition and
concentration of contaminants, type and dose of coagulant and reaction pH (determining the attractive forces
between individual floc components) and the type of flow or energy dissipation (the effect of hydrodynamics).
Flocculation of AOM in Water Treatment 131
Figure 11. Optimal pH ranges for efficient removal of AOM, AOM together with other impurities and
AOM removal efficiencies. Based on Pivokonsky et al. (2012, 2015), Safarikova et al. (2013), Baresova
et al. (2017) and Naceradska et al. (2018).
ACKNOWLEDGMENTS
The research project was funded by the Czech Science Foundation under the Project
No. 18-14445S with institutional support RVO: 67985874. The authors acknowledge the
financial assistance on this project.
REFERENCES
Bache, D. H. & Gregory, R. (2007). Flocs in Water Treatment (1st). London, UK: IWA
Publishing.
Baresova, M., Pivokonsky, M., Novotna, K., Naceradska, J. & Branyik, T. (2017). An
application of cellular organic matter to coagulation of cyanobacterial cells
(Merismopedia tenuissima). Water Research, 122, 70-77.
Bernhardt, H. & Clasen, J. (1991). Flocculation of microorganisms. Journal of Water
Supply: Research and Technology – AQUA, 40(2), 76-87.
Flocculation of AOM in Water Treatment 133
Bernhardt, H., Hoyer, O., Schell, H. & Lüsse, B. (1985). Reaction mechanisms involved
in the influence of algogenic matter on flocculation. Zeitschrift fur Wasser- und
Abwasser-Forschung, 18(1), 18-30.
Bernhardt, H., Lüsse, B. & Hoyer, O. (1986). The addition of calcium to reduce the
impairement of flocculation by algogenic organic matter. Zeitschrift fur Wasser-und
Abwasser-forschung, 19, 219-228.
Bernhardt, H., Shell, H., Hoyer, O. & Lüsse, B. (1991). Influence of algogenic organic
substances on flocculation and filtration. WISA, 1, 41-57.
Bertocchi, C., Navarini, L. & Cesaro, A. (1990). Polysaccharides from Cyanobacteria.
Carbohydrate Polymers, 12, 127-153.
Branyikova, I., Filipenska, M., Urbanova, K., Ruzicka, M. C., Pivokonsky, M. &
Branyik, T. (2018). Physicochemical approach to alkaline flocculation of Chlorella
vulgaris induced by calcium phosphate precipitates. Colloids and Surfaces B:
Biointerfaces, 166, 54-60.
Bubakova, P. & Pivokonsky, M. (2012). The influence of velocity gradient on properties
and filterability of suspension formed during water treatment. Separation and
Purification Technology, 92, 161-167.
Campbell, P. G. C., Twiss, M. R. & Wilkinson, K. J. (1997). Accumulation of natural
organic matter on the surfaces of living cells: implications for the interaction of toxic
solutes with aquatic biota. Canadian Journal of Fisheries and Aquatic Sciences,
54(11), 2543-2554.
Carmichael, W. W. (1992). Cyanobacteria secondary metabolites – the cyanotoxins.
Journal of Applied Bacteriology, 72, 445-459.
Ching, H. W., Tanaka, T. S. & Elimelech, M. (1994). Dynamics of coagulation of kaolin
particles with ferric chloride. Water Research, 28(3), 559–569.
Coral, L. A., Zamyadi, A., Barbeau, B., Bassetti, F. J., Lapolli, F. R. & Prévost, J. (2013).
Oxidation of Microcystis aeruginosa and Anabaena flos-aquae by ozone: Impacts on
cell integrity and chlorination by-product formation. Water Research, 47(9), 2983-
2994.
Creighton, T. E. (1993). Proteins: Structures and Molecular Properties, (2nd). New York,
USA: W.H. Freeman and Company.
Cui, X., Zhou, D., Fan, W., Huo, M., Crittenden, J. C., Yu, Z., Ju, P. & Wang, Y. (2016).
The effectiveness of coagulation for water reclamation from a wastewater treatment
plant that has a long hydraulic and sludge retention times: A case study.
Chemosphere, 157, 224-231.
de Figueiredo, D. R., Azeiteiro, U. M., Esteves, S. M., Gonçalves, F. J. M. & Pereira, M.
J. (2004). Microcystin-producing blooms – a serious global public health issue.
Ecotoxicology and Environmental Safety, 59, 151-163.
Devrimci, H. A., Yuksel, A. M. & Sanin, F. D. (2012). Algal alginate: A potential
coagulant for drinking water treatment. Desalination, 299, 16-21.
134 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Dixon, M. B., Richard, Y., Ho, L., Chow, C. W. K., O’Neill, B. K. & Newcombe, G.
(2011). A coagulation-powdered activated carbon-ultrafiltration-Multiple barrier
approach for removing toxins from two Australian cyanobacterial blooms. Journal of
Hazardous Materials, 186, 1553-1559.
Duan, J. & Gregory, J. (2003). Coagulation by hydrolysing metal salts. Advances in
Colloid and Interface Science, 100-102, 475-502.
Edzwald, J. K. (1995). Principles and applications of dissolved air flotation. Water
Science and Technology, 31(4), 1-23.
Edzwald, J. K. (2010). Dissolved air flotation and me. Water Research, 44(7), 2077-
2106.
Edzwald, J. K. (1993). Coagulation in drinking water treatment: particles, organics and
coagulants. Water Science and Technology, 27(11), 21-35.
Fang, J., Yang, X., Ma, J., Shang, C. & Zhao, Q. (2010). Characterization of algal
organic matter and formation of DBPs from chlor(am)ination. Water Research, 44,
5897-5906.
Filipenska, M., Vasatova, P., Pivokonska, L., Cermakova, L., Naceradska, J. &
Pivokonsky, M. (2018). Influence of COM-peptides/proteins on floc properties
coagulated by metal salts at different shear rate. Langmuir (submitted).
Garzon-Sanabria, A. J., Ramirez-Caballero, S. S., Moss, F. E. P. & Nikolov, Z. L. (2013).
Effect of algogenic organic matter (AOM) and sodium chloride on Nannochloropsis
salina flocculation efficiency. Bioresource Technology, 143, 231-237.
González-Fernández, C. & Ballesteros, M. (2013). Microalgae autoflocculation: An
alternative to high-energy consuming harvesting methods. Journal of Applied
Phycology, 25(4), 991-999.
Gonzalez-Torres, A., Putnam, J., Jefferson, B., Stuetz, R. M. & Henderson, R. K. (2014).
Examination of the physical properties of Microcystis aeruginosa flocs produced on
coagulation with metal salts. Water Research, 60(1), 197-209.
Gonzalez-Torres, A., Rich A. M., Marjo, C. E. & Henderson, R. K. (2017). Evaluation of
biochemical algal floc properties using Reflectance Fourier-Transform Infrared
Imaging. Algal Research, 27, 345-355.
Goring, C. A. I. & Bartholomew, W. V. (1952). Adsorption of mononucleotides, nucleic
acid and nucleoproteins by clays. Soil Science, 74(2), 149-164.
Goslan, E. H., Seigle, C., Purcell, D., Henderson, R., Parsons, S. A., Jefferson, B. &
Judd, S. J. (2017). Carbonaceous and nitrogenous disinfection by-product formation
from algal organic matter. Chemosphere, 170, 1-9.
Greenland, D. J. (1956). The adsorption of sugar by montmorillonite. Journal of Soil
Science, 7(2), 329-334.
Gregor, J. E., Fenton, E., Brokenshire, G., Van Den Brink, P. & O’Sullivan, B. (1996).
Interactions of calcium and aluminium ions with alginate. Water Research, 30(6),
1319-1324.
Flocculation of AOM in Water Treatment 135
Henderson, R., Sharp, E., Jarvis, P. & Jefferson, B. (2006). Identifying the linkage
between particle characteristics and understanding coagulation performance. Water
Science and Technology, 6(1), 31-38.
Henderson, R. K., Baker, A., Parsons, S. A. & Jefferson, B. (2008a). Characterisation of
algogenic organic matter extracted from cyanobacteria, green algae and diatoms.
Water Research, 42(13), 3435-3445.
Henderson, R. K., Parsons, S. A. & Jefferson, B. (2008b). The impact of algal properties
and pre-oxidation on solid–liquid separation of algae. Water Research, 42(8-9),
1827-1845.
Henderson, R. K., Parsons, S. A. & Jefferson, B. (2010). The impact of differing cell and
algogenic organic matter (AOM) characteristics on the coagulation and flotation of
algae. Water Research, 44(12), 3617-3624.
Her, N., Amy, G., Park, H. R. & Song, M. (2004). Characterizing algogenic organic
matter (AOM) and evaluating associated NF membrane fouling. Water Research, 38,
1427-1438.
Hoyer, O., Lusse, B. & Bernhardt, H. (1985). Isolation and characterisation of
extracellular organic matter (EOM) from algae. Zeitschrift fur Wasser-und Abwasser-
Forschung, 18, 76-90.
Hoyer, O., Bernhardt, H. & Lusse, B. (1987). The effect of ozonation on the impairment
of flocculation by algogenic organic matter. Zeitschrift fur Wasser-und Abwasser-
Forschung, 20, 123-131.
Hu, C., Liu, H., Qu, J., Wang, D. & Ru, J. (2006). Coagulation behavior of aluminum
salts in eutrophic water: Significance of Al13 species and pH control. Environmental
Science and Technology, 40, 325-331.
Huang, W. J., Lai, C. H. & Cheng, Y. L. (2007). Evaluation of extracellular products and
mutagenicity in cyanobacteria cultures separated from a eutrophic reservoir. Science
of Total Environment, 377, 214-223.
Chekli, L., Corjon, E., Tabatabai, S. A. A., Naidu, G., Tamburic, B., Park, S. H. & Shon,
H. K. (2017). Performance of titanium salts compared to conventional FeCl3 for the
removal of algal organic matter (AOM) in synthetic seawater: Coagulation
performance, organic fraction removal and floc characteristics. Journal of
Environmental Management, 201, 28-36.
Cheng, Y. S., Zheng, Y., Labavitch, J. M. & Vandergheynst, J. S. (2011). The impact of
cell wall carbohydrate composition on the chitosan flocculation of Chlorella. Process
Biochemistry, 46, 1927-1933.
Jiang, J., Graham, N. J. D. & Harward, C. (1993). Comparison of polyferric sulphate with
othe coagulants for the removal of algae and algae-derived organic matter. Water
Science and Technology, 27(11), 221-230.
136 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Jiao, R., Xu, H., Xu, W., Yang, X. & Wang, D. (2015). Influence of coagulation
mechanisms on the residual aluminum-the roles of coagulant species and MW of
organic matter. Journal of Hazardous Materials, 290, 16-25.
Kim, J. S. & Kang, L. S. (1998). Investigation of coagulation mechanisms with Fe(III)
salt using jar tests and flocculation dynamics. Environmental Engineering Research,
3(1), 11-19.
Knauer, K. & Buffle, J. (2001). Adsorption of fulvic acid on algal surfaces and its effect
on carbon uptake. Journal of Phycology, 37(1), 47-51.
Labille, J., Thomas, F., Milas, M. & Vanhaverbeke, C. (2005). Flocculation of colloidal
clay by bacterial polysaccharides: effect of macromolecule charge and structure.
Journal of Colloid and Interface Science, 284, 149-156.
Laurens, L. M. L., Van Wychen, S., McAllister, J. P., Arrowsmith, S., Dempster, T. A.,
McGowen, J. & Pienkos, P. T. (2014). Strain, biochemistry, and cultivation-
dependent measurement variability of algal biomass composition. Analytical
Biochemistry, 452, 86-95.
Li, L., Gao, N., Deng, Y., Yao, J. & Zhang, K. (2012). Characterization of intracellular &
extracellular algae organic matters (AOM) of Microcystis aeruginosa and formation
of AOM-associated disinfection byproducts and odor & taste compounds. Water
Research, 46, 1233-1240.
Ma, J. & Liu, W. (2002). Effectiveness and mechanism of potassium ferrate(VI)
preoxidation for algae removal by coagulation. Water Research, 36(4), 871-878.
Ma, M., Liu, R., Liu, H., Qu, J. & Jefferson, W. (2012b). Effects and mechanisms of pre-
chlorination on Microcystis aeruginosa removal by alum coagulation: Significance of
the released intracellular organic matter. Separation and Purification Technology, 86,
19-25.
Ma, M., Liu, R., Liu, H. & Qu, J. (2012a). Effect of moderate pre-oxidation on the
removal of Microcystis aeruginosa by KMnO4-Fe(II) process: Significance of the in-
situ formed Fe(III). Water Research, 46(1), 73-81.
Markou, G., Angelidaki, I. & Georgakakis, D. (2012). Microalgal carbohydrates: an
overview of the factors influencing carbohydrates production, and of main
bioconversion technologies for production of biofuels. Applied Microbiology and
Biotechnology, 96, 631-645.
Myat, D. T., Stewart, M. B., Mergen, M., Zhao, O., Orbell, J. D. & Gray, S. (2014).
Experimental and computational investigations of the interactions between model
organic compounds and subsequent membrane fouling. Water Research, 48, 108-118.
Myklestad, S. (1974). Production of carbohydrates by marine planktonic diatoms. I.
Comparison of nine different species in culture. Journal of Experimental Marine
Biology and Ecology, 15(3), 261-274.
Myklestad, S. M. (1995). Release of extracellular products by phytoplankton with special
emphasis on polysaccharides. The Science of the Total Environment, 165, 155-164.
Flocculation of AOM in Water Treatment 137
Naceradska, J., Pivokonsky, M., Pivokonska, L., Baresova, M., Henderson, R. K.,
Zamyadi, A. & Janda, V. (2017). The impact of pre-oxidation with potassium
permanganate on cyanobacterial organic matter removal by coagulation. Water
Research, 114, 42-49.
Naceradska, J., Novotna, K., Cermakova, L., Cajthaml, T. & Pivokonsky, M. (2018).
Investigating the coagulation of non-proteinaceous algal organic matter: optimization
of coagulation performance and identification of removal mechanisms. Journal of
Environmental Sciences (submitted).
Ngo, H. H., Vigneswaran, S. & Dharmappa, H. B. (1995). Optimization of direct
filtration: Experiments and mathematical models. Environmental Technology, 16(1),
55-63.
Nguyen, M. L., Westerhoff, P. E., Baker, L., Hu, Q., Esparza-Soto, M. & Sommerfeld,
M. (2005). Characteristics and Reactivity of Algae-Produced Dissolved Organic
Carbon. Journal of Environmental Engineering, 131(11), 1547-1582.
Nicolaus, B., Panico, A., Lama, L., Romano, I., Manca, M. C., De Giulio, A. &
Gambacorta, A. (1999). Chemical composition and production of exopolysaccharides
from representative members of heterocystous and non-heterocystous cyanobacteria.
Phytochemistry, 52, 639-647.
Paerl, H. W., Fulton, III, R. S., Moisander, P. & Dyble, J. (2001). Harmful Freshwater
Algal Blooms, With an Emphasis on Cyanobacteria. The Scientific World, 1, 76-113.
Paralkar, A. & Edzwald, J. K. (1996). Effect of ozone on EOM and coagulation. Journal
of the American Water Works Association, 88(4), 143-154.
Parfit, R. L. & Greenland, D. J. (1970). Adsorption of polysaccharides by
montmorillonite. Soil Science Society of America Journal, 34(6), 862-866.
Parker, D. S., Kaufman, W. J. & Jenkins, D. (1972). Floc breakup in turbulent
flocculation processes. Journal of the Sanitary Engineering Division, 98(1), 79-99.
Pearson, L. A., Dittmann, E., Mazmouz, R., Ongley, S. E., D’Agostino, P. M. & Neilan,
B. A. (2016). The genetics, biosynthesis and regulation of toxic specialized
metabolites of cyanobacteria. Harmful Algae, 54, 98-111.
Pinck, L. A. (1962). Adsorption of proteins, enzymes and antibiotics by montmorillonite.
9th National conference on clays and minerals: 520-529.
Pivokonsky, M., Kloucek, O. & Pivokonska, L. (2006). Evaluation of the production,
composition and aluminium and iron complexation of algogenic organic matter.
Water Research, 40, 3045-3052.
Pivokonsky, M., Naceradska, J., Brabenec, T., Novotna, K., Baresova, M. & Janda, V.
(2015). The impact of interactions between algal organic matter and humic
substances on coagulation. Water Research, 84, 278-285.
Pivokonsky, M., Naceradska, J., Kopecka, I., Baresova, M., Jefferson, B., Li, X. &
Henderson, R. K. (2016). The impact of algogenic organic matter on water treatment
138 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Vandamme, D., Beuckels, A., Vadelius, E., Depraetere, O., Noppe, W., Dutta, A.,
Foubert, I., Laurens, L. & Muylaert, K. (2016). Inhibition of alkaline flocculation by
algal organic matter for Chlorella vulgaris. Water Research, 88, 301-307.
Vandamme, D., Foubert, I., Fraeye, I. & Muylaert, K. (2012). Influence of organic matter
generated by Chlorella vulgaris on five different modes of flocculation. Bioresource
Technology, 124, 508-511.
Vandamme, D., Muylaert, K., Fraeye, I. & Foubert, I. (2014). Floc characteristics of
Chlorella vulgaris: Influence of flocculation mode and presence of organic matter.
Bioresource Technology, 151, 383-387.
Vigneault, B., Percot, A., Lafleur, M. & Campbell, P. G. C. (2000). Permeability changes
in model and phytoplankton membranes in the presence of aquatic humic substances.
Environmental Science and Technology, 34(18), 3907-3913.
Villacorte, L. O., Ekowati, Y., Neu, T. R., Kleijn, J. M., Winters, H., Amy, G., Schippers,
J. C. & Kennedy, M. D. (2015). Characterisation of algal organic matter produced by
bloom-forming marine and freshwater algae. Water Research, 73, 216-230.
Vlaški, A., van Breemen, A. N. & Alaerts, G. J. (1997). The role of particle size and
density in dissolved air flotation and sedimentation. Water Science and Technology,
36(4), 177-189.
Wang, L., Qiao, J., Hu, Y., Wang, L., Zhang, L., Zhou, Q. & Gao, N. (2013). Pre-
oxidation with KMnO4 changes extra-cellular organic matter's secretion
characteristics to improve algal removal by coagulation with a low dosage of
polyaluminium chloride. Journal of Environmental Sciences, 25(3), 452-459.
Widrig, D. L., Gray, K. A. & McAuliffe, K. S. (1996). Removal of algal-derived organic
material by preozonation and coagulation: Monitoring changes in organic quality by
pyrolysis-GC-MS. Water Research, 30(11), 2621-2632.
Wu, Z., Zhu, Y., Huang, W., Zhang, C., Li, T., Zhang, Y. & Li, A. (2012). Evaluation of
flocculation induced by pH increase for harvesting microalgae and reuse of
flocculated medium. Bioresource Technology, 110, 496-502.
Wyatt, N. B., Gloe, L. M., Brady, P. V., Hewson, J. C., Grillet, A. M., Hankins, M. G. &
Pohl, P. I. (2012). Critical Conditions for Ferric Chloride-Induced Flocculation of
Freshwater Algae. Biotechnology and Bioengineering, 109(2), 493-501.
Zhang, X., Amendola, P., Hewson, J. C., Sommerfeld, M. & Hu, Q. (2012). Influence of
growth phase on harvesting of Chlorella zofingiensis by dissolved air flotation.
Bioresource Technology, 116, 477-484.
Zhang, X., Fan, L. & Roddick, F. A. (2014). Feedwater coagulation to mitigate the
fouling of a ceramic MF membrane caused by soluble algal organic matter.
Separation and Purification Technology, 133, 221-226.
Zhang, X. J., Chen, C., Ding, J. Q., Hou, A., Li, Y., Niu, Z. B., Su, X. Y., Xu, Y. J. &
Laws, E. A. (2010). The 2007 water crisis in Wuxi, China: Analysis of the origin.
Journal of Hazardous Materials, 182, 130-135.
140 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
BIOGRAPHICAL SKETCH
Martin Pivokonský
Professional Appointments:
Naceradska, J., Novotna, K., Cermakova, L., Cajthaml, T. & Pivokonsky, M. (2018).
Investigating the coagulation of non-proteinaceous algal organic matter: optimization
of coagulation performance and identification of removal mechanisms. Journal of
Environmental Sciences (submitted).
Filipenska, M., Vasatova, P., Pivokonska, L., Cermakova, L., Naceradska, J. &
Pivokonsky, M. (2018). Influence of COM-peptides/proteins on floc properties
coagulated by metal salts at different shear rate. Langmuir (submitted).
Branyikova, I., Filipenska, M., Urbanova, K., Ruzicka, M. C., Pivokonsky, M. &
Branyik, T. (2018). Physicochemical approach to alkaline flocculation of Chlorella
vulgaris induced by calcium phosphate precipitates. Colloids and Surfaces B:
Biointerfaces, 166, 54-60.
Angst, G., Mueller, C.W., Angst, Š., Pivokonský, M., Franklin, J., Stahl, P.D. & Frouz, J.
(2018). Fast accrual of C and N in soil organic matter fractions following post-
mining reclamation across the USA. Journal of Environmental Management, 209,
216-226.
Cermakova, L., Kopecka, I., Pivokonsky, M., Pivokonska, L. & Janda, V. (2017).
Removal of cyanobacterial amino acids in water treatment by activated carbon
adsorption. Separation and Purification Technology, 173, 330-338.
Baresova, M., Pivokonsky, M., Novotna, K., Naceradska, J. & Branyik, T. (2017). An
application of cellular organic matter to coagulation of cyanobacterial cells
(Merismopedia tenuissima). Water Research, 122, 70-77.
Naceradska, J., Pivokonsky, M., Pivokonska, L., Baresova, M., Henderson, R. K.,
Zamyadi, A. & Janda, V. (2017). The impact of pre-oxidation with potassium
142 Martin Pivokonský, Jana Načeradská, Kateřina Novotná et al.
Chapter 6
ABSTRACT
The leather tanning industry uses a large amount of toxic substances and water.
Thus, it generates effluents with high polluting load. Due to the large amount of effluent
generated by tanneries and the difficulty in their treatment, several studies have been
proposed to minimize the environmental impacts caused by inappropriate disposal of this
effluent. This study investigated the performance of natural coagulants Tanin compared
to chemical coagulants aluminium sulphate and ferric chloride commonly used in the
treatment of raw wastewater from tannery, by means of the physicochemical processes of
coagulation, flocculation and sedimentation. Using jar tests methods, different
concentrations for the coagulants (chemical and natural) were applied to the concerned
effluent and it was assessed their efficiency in removing certain parameters such as
apparent color COD, and turbity besides the behavior of electrical conductivity and pH.
The results showed that for pH and electrical conductivity parameters, there was no
significant variation after application of coagulant in relation to the raw effluent. By
comparing the coagulants for colour parameter it was observed that the natural coagulant
144 E. R. Pereira, G. da Silva Souza and J. D. P. Theodoro
tanin had greater efficiency compared to the chemical ones, reaching 56.9% of removal,
however, the aluminium sulphate was more efficient at removing COD, with 95% of
removal. Therefore, it is noticed that the chemical coagulant aluminium sulphate was
more successful for the majority of the analysed parameters in comparison to other
coagulants.
1. INTRODUCTION
2. METHODOLOGY
The effluent used in this research came from a leather tanning industry, industry
located in the city of Ibiporã – PR, Brazil. The effluent was collected in the treatment
system of the still raw industry, before the application of the coagulant used by the
industry. After being collected, the effluent was sent for testing and experimental analysis
at the Sanitation Laboratory of the Federal Technological University of Paraná - Campus
Londrina. A pre-test was performed first, aiming at defining the most appropriate
concentrations of the chemical and natural coagulants used. Thus, the following
concentrations were obtained: for Tannin C1 3mgL-1, C2 6mgL-1, C3 9mgL-1; for
aluminum sulfate C1 15mgL-1, C2 20mgL-1, C3 25mgL-1; and for ferric chloride C1
15mgL-1, C2 21mgL-1, C3 28mgL-1.
To prepare the coagulants 100 g of each coagulant were weighed and diluted in 1 L
of distilled water using a volumetric flask. After the concentrations of each coagulant
were defined in the pre-test, these concentrations were added to the effluent in a
coagulation/flocculation/sedimentation test using the Jar-test equipment.
The methodology for the test in the Jar Test was adapted from Theodoro (2012). The
parameters analyzed are described in Table 1 and are in accordance with APHA (2012).
The statistical model used to perform the analysis of the values resulting from the
four trials was the completely randomized design, in a factorial scheme. The statistical
model considered the effect of the interaction between Coagulant and Concentration
factors in addition to the effect of the Coagulant factor (in three levels) and the factor
Concentration (in three levels). Thus, the multiplicative statistical model is given by
equation 1:
On what:
Yijk = observation corresponding to the kth experimental unit submitted to the i-th
level of the Coagulant factor and the jth level of the factor Concentration;
: overall average common to all observations;
i : effect of the ith level of the coagulant factor;
j : effect of the jth level of the factor Concentration;
ij: is the interaction effect between the i-th level of the Coagulant factor and
the j-th level of the factor Concentration;
eijk: component of the random error associated with observation Yijk.
For the analysis of variance and comparison of the means of the three variables, a
level of significance of 10% was considered. The results of the analysis of variance were
grouped in Table 2 as proposed by Martins (2006).
3. RESULTS
The Table 3 shows the values with pH variation during sedimentation times.
For the coagulant Tannin, it was observed that there was not a very significant
variation in relation to the initial pH (12.97), and the greatest variation occurred in the
time of 30 minutes, when the lowest concentration of 300 mgL-1 reached the value of
12.90. The three concentrations obtained similar values and a little below the initial one,
varying from 12.90 to 12.92. The pH can be adjusted for an application of the Aluminum
Sulphate, in this way, the pH must be improved to be removed.
When analyzing the coagulant Ferric Chloride, it was observed that there was the
greatest reduction in the pH value with the highest coagulant concentration, reaching
12.53. After this, it can be observed that there was an increase in pH values for all
concentrations.
When comparing the data obtained with CONAMA Resolution 430 (CONAMA,
2011), it is observed that even after the treatment done in the effluent, it does not fit into
Comparison of Natural Coagulant and Chemistry in Tanning … 147
the standards of release standard in relation to pH, since the Resolution requires that the
pH value is maintained between 5 and 9 to be released into a receiving body. In this case,
the pH of the tannery effluent should be corrected so that its value becomes less alkaline
and falls within the parameters required by the law for the pH parameter. There was no
significant variation in the pH values for tannery effluent with both natural and chemical
coagulants.
Figure 1 shows the behavior of the coagulants adopted in the work for the percentage
of removal vesus time of the apparent color. The effluent collected in crude form had a
color value of 96.800 mgPtCo L-1.
Figure 1a shows that for the tannin all the concentrations used reached an
apparent color removal greater than 56%. When analyzing the efficiency of
the Coagulant Aluminum Sulphate (Figure 1b) it is observed that the largest removal
was 36%. It is noted that the removal of the chemical coagulant Aluminum Sulphate
was lower than the percentage of removal of the natural coagulant
Tanino which demonstrates the viability of the organic coagulant for the effluent in
question.
In Figure 1c when looking at the Ferric Chloride as a coagulant agent in the removal
of apparent color, it was observed that it is not a promising agent in the removal of color,
because there was an increase in the parameter in all the concentrations used.
148 E. R. Pereira, G. da Silva Souza and J. D. P. Theodoro
(a)
(b)
(c)
Figure 1. Percentages of removal of the apparent color parameter as a function of time for the
coagulants Tanino (a), Aluminum Sulphate (b) and Ferric Chloride (c).
Comparison of Natural Coagulant and Chemistry in Tanning … 149
In a study carried out by Vaz (2010), a similar result was obtained concluding the
author that when the ferric chloride is added with some excess to the effluent, part of it
does not participate in the coagulation/flocculation reaction, being this in solution having
an increase of the values of the apparent color parameter.
Statistical analysis showed a P-value of 0.841 > 0.1, indicating that the interaction
between the Coagulant and Concentration factors is not significant, at 10% significance
and it is also noted that the Concentration is not significant (value- P = 0.928 > 0.1), but
the effect of Coagulant was significant (P-value = 0.000 < 0.1). The same occurred for
the second, third and fourth sedimentation times, thus, mean comparisons were
performed only for Coagulant and are presented in Table 4, using the Tukey test.
The organic coagulant Tanino presented in all times of sedimentation statistically
different averages and values statistically smaller than the other coagulants, thus
confirming its greater efficiency in the removal of apparent color in comparison with the
other coagulants.
The Table 5 shows the behavior of the electrical conductivity during the
sedimentation process in the assays for each coagulant. The effluent of tannery had an
electrical conductivity of 21.84 mScm-1 before the treatment done in this work.
For the coagulant Tannin, it is possible to observe that in all the concentrations used
in the test, an increase occurred in the electrical conductivity, where in the time of
sedimentation of 40 minutes the highest value was observed, reaching this 22.57 mScm-1
using the lowest concentration of the coagulant.
150 E. R. Pereira, G. da Silva Souza and J. D. P. Theodoro
When analyzing the results of the electrical conductivity for the Coagulant
Aluminum Sulphate, it can be said that there was an insignificant variation. For the
coagulant Ferric Chloride, it was observed that there was an increase in the electric
conductivity for the highest concentration of coagulant used. At the lower concentration,
there was a decrease in conductivity values. In general, it is noticed that there was not a
significant change in the electric conductivity after the coagulants application and, in this
way, in relation to the parameter the coagulants do not have any variation of the same.
Figure 2 shows the behavior of fixed and volatile solids in the treatment of tannery
effluent. The crude effluent values for fixed solids and for volatile solids are 60,430 and
38,200 mgL-1, respectively. For the natural coagulant Tanino, it is observed in Figure 2
that the intermediate concentration of 6 mgL-1 was the one that best behaved in the
removal of the solids, reaching a 48% removal for the fixed solids and 59% for the
volatile solids. By observing the behavior of the chemical coagulant Aluminum Sulphate
in the removal of solids, it was observed that it was efficient reaching values above 70%
in all concentrations. The highest removal of fixed solids was 75% and occurred at the
intermediate concentration of 20 mgL-1, the highest removal of volatile solids was 83%
and occurred at the concentration of 28 mgL-1.
In relation to the chemical coagulant Ferric Chloride, as well as the natural coagulant
Tanino, the greatest removal of the solids occurred for the intermediate concentration. At
the concentration of 21 mgL-1 there was a 51% removal of fixed solids and 59% of
volatile solids, indicating values similar to Tannin and a similar behavior for solids
Comparison of Natural Coagulant and Chemistry in Tanning … 151
removal, however, the chemical coagulant uses a higher concentration to obtain this
removal.
Figure 2. Solids series for the coagulants Tannin (a), Aluminum Sulphate (b)
and Ferric Chloride (c), in percentage values. TC1 = Tannin 3 mgL-1; TC2 = Tannin
6 mgL-1; TC3 = Tannin 9 mgL-1; SC1 = Aluminum Sulfate 15 mgL-1; SC2 = Aluminum Sulphate 20
mgL-1; SC3 = Aluminum Sulphate 25 mgL-1; CCl = Ferric Chloride 15 mgL-1; CC2 = Ferric Chloride
21 mgL-1; CC3 = Ferric Chloride 28 mgL-1.
(a) (b)
(c)
Figure 3. Removal percentage of the Chemical Oxygen Demand parameter for the coagulants Tanino
(a), Aluminum Sulphate (b), and Ferric Chloride (c).
152 E. R. Pereira, G. da Silva Souza and J. D. P. Theodoro
Figure 3 shows the percentage of COD removal obtained for the coagulants
used in the assay. The raw effluent collected at the company had a COD value of
79.850 mgL-1.
For the parameter COD, it is possible to infer that the best results were obtained with
the chemical coagulants, reaching values higher than 90% of COD removal, since no
natural coagulant did not obtain values higher than 80%. Although the coagulants,
especially the chemical ones, present great removal, the effluent still could not be
released in the water body, since it does not meet the standards established by CEMA no.
70/2009 (CEMA, 2009) in annex 7, exceeding the limit value for COD, which is 350
mgL-1 for tanneries.
CONCLUSION
Based on this research, it was verified that the chemical coagulant aluminum sulphate
presented better efficiency in relation to the coagulant tannin. Therefore, the application
of aluminum sulphate would be more feasible with regard to the physical-chemical
treatment of the tanning effluent.
REFERENCES
APHA – American Public Health Association. Standard Methods for the Examination of
Water and Wastewater. 22 ed. Washington, 2012.
Bayer, V. Study of the extraction of hexavalent chromium, by the technique of surfactant
liquid membranes, aiming the treatment of liquid effluents of Tanneries. 2005. 126f.
Dissertation (Master’s Degree in Chemical Engineering - Post-Graduation Program
in Chemical Engineering, Federal University of Minas Gerais, Belo Horizonte, 2012.
Brazil. National Environment Council - Conama. Resolution No. 357, dated March 17,
2005. Provides for the classification of water bodies and environmental guidelines for
their classification, as well as establishing the conditions and standards for effluent
discharge, and other measures. Official Gazette of the Union. Executive Power,
Brasilia, DF, March 18. 2005. Available at: <http://www.mma.gov.br/port/
conama/res/res05/res35705.pdf>. Accessed on: May 25th. 2014.
Brazil. National Environment Council - Conama. Resolution no. 430, dated May 13,
2011. Provides for conditions and standards for the discharge of effluents,
complements and amends Resolution 357 of March 17, 2005, of the National Council
for the Environment - CONAMA. Official Journal of the Union. Executive Branch,
Comparison of Natural Coagulant and Chemistry in Tanning … 153
Chapter 7
ABSTRACT
The proposal of this paper was to evaluate through physical and chemical parameters
the efficiency of the coagulation/flocculation/ sedimentation/filtration processes using
organic coagulant (Moringa oleifera) in the treatment of water from a lentic system
(Igapó II Lake), located in Londrina – Paraná - Brasil. It was made the water collection at
two points of the lake, in the entrance (point 1) and at the exit (point 2). The treatment
occurred using the jar-test equipment with the same conditions of fast mixing and slow
mixing used in the Water Treatment Station (ETA). The sedimentation time was 30
minutes. The granulometry of the sand used in the filters was the same, in the range of
0.600 to 0.850 mm. The concentrations of saline solution of the organic coagulant to be
applied were: C1 = 3 mg.L-1, C2 = 6 mg.L-1 and C3 = 9 mg.L-1. In addition, it was
necessary to correct the pH of the samples in order to evaluate the behavior of the organic
coagulant in different pH levels. After the assays, a study was carried out on the removal
of turbidity and apparent color after the coagulation / flocculation / sedimentation
processes and after the filtration process. It was verified for the parameters of turbidity
and of apparent color, the concentration C2 (6 mg.L-1) and the pH around 7 (neutral)
presented the highest removal efficiency, being it above 99%. For the pH, all
156 J. C. Belisário Junior, E. R. Pereira and J. D. Peruço Theodoro
concentrations did not present great variations when compared to the crude sample, in
addition the results prove the efficiency of the coagulant in different pH levels. Only
assay 2 in the entrance point, did not meet the maximum values allowed by the potability
ordinance 2,914/11 of the ministry of health for parameters of apparent color and of
turbidity, all other assays are in accordance. The results proved the efficiency of the
organic coagulant extracted from Moringa oleifera seeds for the treatment of water in
lentic systems.
INTRODUCTION
The fresh water is divided into lentic ecosystems, lotic ecosystems and flooded
ecosystem. The lentic systems are characterized by low flow or absence of hydrological
flow, being represented by lakes, ponds and tanks. The lotic systems are characterized by
the presence of water flow, such as rivers, streams and springs, on the other hand flooded
systems, have water levels floating up and down during the year, being represented by
swamps and marshes (Soares, 2005).
According to the World Health Organization, about 80% of all diseases that spread in
developing countries come from poor water. Diseases such as typhoid fever, cholera,
diarrhea among others (Richter; Netto, 1991).
The treatment of water aims to reduce impurities and/or correct some aspects of raw
water with the possibility of reuse, or make it drinkable (Freitas, 2011).
Vanacor (2005) states that in addition to making it drinkable, water treatment must
meet the community’s need to have good quality water from a chemical, physical and
bacteriological point of view, and may also serve hygienic purposes (removal of bacteria,
viruses, microorganisms, algae, undesirable substances), aesthetics (color correction,
turbidity, odor and taste) and economic (reduction of corrosivity, hardness, iron,
manganese, etc.).
Coagulation is one of the fundamental processes in surface water treatment systems
for public supply purposes, as it is responsible for water clarification, removal of most
heavy metals, and chemical and microbiological agents (Macedo, 2007).
Coagulation and flocculation are practically simultaneous and interdependent, to the
point that they can be considered a single step: coagulation / flocculation. Coagulation /
flocculation is the process in which the particles agglutinate into small masses (flakes)
with a specific weight higher than that of water. The transport of these particles into the
liquid causes contact, establishing bridges between the particles and forming a three-
dimensional mesh of porous clots (Vanacor, 2005).
Moringa oleifera Seed Use in Salina Solution … 157
METHODOLOGY
The water samples were taken from a lentic system from Lake Igapó II, located in the
city of Londrina, state of Paraná, Brazil. Water samples of entry (Point 1) and exit (Point
2) were used. Figure 1 spatially represents the location of the collection points of each
sample.
Sampling was carried out at the entrance (Point 1) and exit (Point 2) of Lake Igapó II,
highlighted according to Figure 1 and conditioned in plastic gallons of 50 liters each,
previously sanitized.
158 J. C. Belisário Junior, E. R. Pereira and J. D. Peruço Theodoro
Figure 1. Scheme of the satellite view of the location of Lake Igapó II.
For the preparation of the Moringa oleifera organic coagulant solution, first the
Moringa oleifera seed was peeled and 10g of it was weighed. Also weighed 58.44 g of
NaCl and added to 1 L of distilled water, the solution was stirred well for the salt to
dissolve completely, the salt solution being 1M NaCl. Finally, the 10 g of the seeds were
ground together with the saline solution in the blender. The solution was then filtered
using a cloth strainer and the Moringa oleifera coagulant saline at a concentration of 10
g.L-1 was obtained.
To adjust the pH values to make it acidic, the solution of hydrochloric acid (HCl),
concentration 1 M was used. It was necessary to add 20 mL of HCl to the volume used in
order to achieve the acidic pH set. Meanwhile, the sodium hydroxide solution (NaOH), at
1 M concentration, was used for the basic pH sample. It was necessary to add 20 mL of
the basic solution to achieve the set basic pH.
Fast Mixing Slow Mixing Slow Mixing Slow Mixing Slow Mixing
Gradient (s-1) 540 90 52 40 30
Time (min) 00:10 02:00 02:35 02:40 05:40
Source. (Trevisan et al. (2014); Higashi (2016); Goes et al., (2017))
Moringa oleifera Seed Use in Salina Solution … 159
The final pH values applied in the samples are shown in Table 1. For the experiment,
the dosages found for coagulant at both points were varied following the concentrations 3
mg.L-1, 6 mg.L-1 and 9 mg.L-1.
The assays were carried out in the Jar-test equipment of six tests with regulator of
rotation of the mixer rods. Placing in each compartment of the 2L water equipment with
different pH and its respective concentration of extraction solution and natural coagulant.
After it was held equipment rotation process at different rates (Table 2) for that to happen
the coagulation and flocculation processes.
For the filtration assay an iron structure adapted to fix the sand filters below the jar-
test was used so that the water exited the jar-test directly into the filters. The filter beds
are made of Polyethylene Tereftalo (PET) of approximately 10 cm of internal diameter
configuring a model of fixed bed with downward flow, with six columns in parallel, as
observed in Figure 2.
It was recommended by Di Bernardo et al., (2003) the tubes are 25 cm long with 15
cm filled by sand, 3 cm filled with crushed stone and 2 cm filled with cotton. The
granulometric values of the sand used in the 6 filters were the same, in the range of 0.600
to 0.850 mm. The slope used at the exit of the test jar was 70 ° for 2 minutes, followed by
an additional 2 minutes at an angle of 60 ° and a further 2 minutes at 50 °. A model of the
filter layers is shown in Figure 3.
The water to be treated entered through the top of the column and was withdrawn
from the bottom. The grains of the filter bed were retained by a coffee filter together with
the cotton and crust layer present at the lower end of the column. After this process, all
parameters were measured.
It is important to say that the assays were performed in duplicate for each pH
value and concentration. For all variations of Moringa oleifera and pH concentrations
the parameters of apparent color, turbidity and pH were determined (Table 3)
according to the Standard Methods of Examination of Water and Wastewater (APHA,
2012).
Parameter Pattern
Apparent color 15 mgPtCo.L-1
Turbidity 5 NTU
pH 6.0 a 9.0
Source: Ministry of Health, Ordinance 2914/11 PRC September 20, 2017 and consilidation ordinance no5, of
September 28, 2018.
Moringa oleifera Seed Use in Salina Solution … 161
The results were evaluated and compared with the standard allowed by ordinance
2,914/11 updated to consolidation ordinance no5, of September 28, 2018 both from the
ministry of health (Table 4) and can indicate if the experimental procedure was
satisfactory.
In this work, statistical analysis was used with two factors (independent
variables), the concentration of the Moringa oleifera coagulant, and pH. The assays
had a repeat, thus, performed in duplicate in Table 5 it is possible to verify these
assay values.
RESULTS
Figures 4, 5 and 6 show the removals of the apparent color parameter for the
coagulation / flocculation / sedimentation processes (Figure 4), for the filtration process
(Figure 5) and for the complete process, coagulation / flocculation / sedimentation /
filtration (Figure 6) with the use of Moringa oleifera in different values of concentration
and pH.
Analyzing the presented results, it can be concluded that both the coagulation /
flocculation / sedimentation processes (Figure 4) and the filtration process (Figure 5) are
162 J. C. Belisário Junior, E. R. Pereira and J. D. Peruço Theodoro
of great importance for the removal of color in the water treatment, both reached removal
values above 80%. The processes together (complete treatment) demonstrated in almost
all tests percent removal of color close to 100% as can be seen in Figure 6.
Figure 4. Removal of the apparent color parameter after the coagulation / flocculation / sedimentation
processes.
Figure 5. Removal of the apparent color parameter after the filtration process.
Figure 6. Removal of the apparent color parameter after the coagulation / flocculation / sedimentation /
filtration process.
In Figure 4 it is possible to verify the assay 2 as the assays with the lowest percentage
of color removal, that is, a concentration of 3 mg.L-1 and a neutral pH of 6.8. However,
Moringa oleifera Seed Use in Salina Solution … 163
its replica at the same point was the one with the highest removal value, besides that in
point 2 the values of color removal with these same variables were also satisfactory.
Also in Figure 4, assays 9 and 18 in point 2 were the assays with the lowest
percentage of color removal at this point, these assays were replicates of each other, that
is, concentration of 18 mg.L-1 and basic pH equal to 10.8. Comparing with the assays of
point 1 it is verified that for point 1 these tests were also the ones with lower percentages
of removal. With this we conclude that for color removal this combination was the worst.
The other assays are ranging from 81% to 91% color removal after coagulation /
flocculation / sedimentation processes.
In the filtration process (Figure 5) it is also possible to verify a small variation of the
percentages of color removal between 85% and 97%, except test 11 of item 1, which had
removal value of 70%. Its replica at the same point also had below-average value, so it is
understood that for such pH value and concentration the filtration process is slightly
affected.
In general, very satisfactory results were obtained, as can be seen in Figure 6, the
complete treatment provided a reduction of the parameter apparent color of almost 100%
in all the tests, it can be concluded that the processes are complementary and very
efficient for the removal of color in the treatment of water. According to studies by
Okuda et al., (1999), which demonstrate that the percentage of color removal using both
the seed and the aqueous extract of Moringa oleifera was 80 to 99%.
It is also possible to verify that the coagulant Moringa oleifera can be effective in a
wide range of pH and concentration in this case, without significantly altering its results
of color removal with the variation of these variables. According to studies conducted by
Vaz (2009), Moringa oleifera can be used in a pH range between 4.0 and 12.0.
Comparing the values obtained with those required by ordinance 2,914/2011 of the
ministry of health 2017 and consilidation ordinance no5, of September 28, 2018, only one
assay did not meet the VMP of apparent color, assay 2 of point 1. All others are in
accordance with the amount required for the use of water for urban supply.
Figure 7. Removal of the turbidity parameter after the coagulation / flocculation / sedimentation
process.
164 J. C. Belisário Junior, E. R. Pereira and J. D. Peruço Theodoro
Figure 9. Removal of the turbidity parameter after the coagulation / flocculation / sedimentation /
filtration process.
Figures 7, 8 and 9 show the removals of the turbidity parameter for the coagulation /
flocculation / sedimentation processes (Figure 7), for the filtration process (Figure 8) and
for the complete process, coagulation / flocculation / sedimentation / filtration (Figure 9)
with the use of Moringa oleifera in different values of concentration and pH.
Analyzing the results presented in Figures 7 and 8, it is possible to verify that both
the coagulation / flocculation / sedimentation processes and the filtration process had a
high percentage of turbidity parameter removal, around 80% in Figure 7 and around 90%
in Figure 8. With this it is possible to understand that the processes together are of
extreme importance for the removal of turbidity in the treatment of water. Confirming the
data presented in Figure 9 where the processes together (complete treatment)
demonstrated in almost all the tests percentage of turbidity removal close to 100%.
This confirms studies carried out by Higashi (2015), where after the treatment of
Lake Igapó II water through the coagulation / flocculation / sedimentation processes with
the use of organic coagulant Moringa oleifera obtained removal results above 80%.
All other assays are ranging from 77% to 88% turbidity removal after coagulation /
flocculation / sedimentation processes.
Moringa oleifera Seed Use in Salina Solution … 165
CONCLUSION
The results of the present paper showed that the use of organic coagulant from the
Moringa oleifera seed used in a saline solution (NaCl, 1M) was very efficient in the
removal of the parameters apparent color and turbidity. For the parameters of apparent
color and turbidity, all the assays, at the end of the complete treatment, showed almost
100% removals.
REFERENCES
APHA – American Public Health Association. Standard Methods for the Examination of
Water and Wastewater. 22 ed. Washington, 2012.
Arantes, C. C. et al. Different forms of Moringa oleifera seed in water treatment.
Brazilian Journal of Agricultural and Environmental Engineering. v.19, n.3, p.266-
272, 2015.
Brazil, Ministry of Health. 2,914, dated December 12, 2011. Available at:
<http://bvsms.saude.gov.br/bvs/saudelegis/gm/2011/prt2914_12_12_2011.html>.
Accessed on: January, 2018.
Brazil, Ministry of Health. Consolidation Ordinance No. 5 of September 28, 2017.
Available at: <http://bvsms.saude.gov.br/bvs/saudelegis/gm/2017/prc0005_03_10_
2017.html> Accessed: April 11 2018.
166 J. C. Belisário Junior, E. R. Pereira and J. D. Peruço Theodoro
Di Bernardo, L.; Mendes, C. G. N.; Brandão, C. C. S.; Sens, M. L.; Pádua, V. L. Water
Treatment for Direct Filtration. Di Bernardo, Luiz (coordinator) - Rio de Janeiro:
ABES, RiMa, 2003. Project PROSAB 498 p.
Di Bernardo, L.; Dantas, A.D. B. Methods and Techniques of Water Treatment. 2nd
Edition. São Carlos: Rima, 2005.
Freitas, D. B. Study of improvements in the chlorination systems of the water supply in
the corsan by the installation of chlorine evaporators and gas scrubbers. Porto
Alegre. 2011. Diploma in Chemical Engineering. Federal University of Rio Grande
do Sul.
Goes, H. H. D.; Souza, R. C. P.; Melo, J. M.; Theodoro, J. D. P.; Study of the application
of Opuntia cochenillifera cactus in water treatment. Encyclopedia Biosphere,
Knowing Scientific Center - Goiânia, v.14 n.25; P. 554-563. 2017.
Higashi, V. Y.; Theodoro, J. D. P.; Pereira, E. R.; Theodoro, P. S.; Use Of Chemical
Coagulants (Ferric Chloride) and Organic (Moringa Oleifera) In Treatment Of
Waters Derived From The Lentic System. Technical Scientific Congress of
Engineering and Agronomy – CONTECC’2016, August 29 to September 2, 2016 -
Foz do Iguaçu, Brazil. Available at: <https://www.tratamentodeagua.com.br/wp-
content/uploads/2016/11/Uso-de-coagulantes-qu%C3%ADmico-cloreto-ferrico-e-
org%C3%A2nico-moringa-oleifera-in-treatment-of-water-from-system-
l%C3%AAntico.pdf>. Accessed in Feb. 2018.
Macêdo, J. A. B. Waters & Waters: 3 ed. Belo Horizonte. MG: CRQ-MG, 2007. 1048p.
Ndabigengesere, A.; Narasiah, K. S.; Talbot, B. G. Active agents and mechanism of
coagulation of turbid water using Moringa oleifera. Water Research. V.29, p.703-
710, 1995.
Okuda, T. et al. Improvement of extraction method of coagulation active components
from Moringa oleifera seed. Water Res. V.33, n.15, p.3373-3378. 1999.
Pichler, T., Young, K.; Alcantar, N. Eliminating turbidity in drinking water using the
mucilage of a common cactus. Water Science and Technology: Water Supply, v. 12,
n.2, 179-186, 2012.
Richter, C. A.; Netto, J. M. A. Water treatment: Upgraded technology. São Paulo:
BLUCHER, 1991. 332p.
Soares, D. H. G. Ecosystems in the ecological context. Pernambuco. Monography
(Undergraduate). Center for Higher Education Arcoverde, 2005.
Theodoro, J. D. P.; Lenz, G. F.; Zara, R. F.; Bergamasco, R.. Coagulants and Natural
Polymers: Perspectives for the Treatment of Water. Plastic and Polymer Technology
(PAPT), v. 2, Issue 3, September 2013.
Trevisan, Thales S. Coagulant Tanfloc SG as an alternative to the use of chemical
coagulants in the treatment of water in ETA Cafezal. 2014. 106 f. Course Completion
Work (Undergraduate) - Superior Course in Environmental Engineering. Federal
Technological University of Paraná, Londrina, 2014.
Moringa oleifera Seed Use in Salina Solution … 167
Chapter 8
ABSTRACT
*
Corresponding Author Email: [email protected].
170 E. A. López-Maldonado and M. T. Oropeza-Guzmán
chitosan the treated water has the following values of BOD5 = 30 mg O2/L, fats and oil =
1 mg/L, turbidity = 2 FAU, TSS = 2 mg/L and COD = 50 mg/L.
1. INTRODUCTION
One of the issues of greatest scientific and technological interest is the quality and
use of water, derived from the problems of scarcity and water contamination. The
interaction of water with the environment is influenced by the water quality that both
society and the industrial sector confer on water.
Each type of industry has a particular interest in the care of water quality and its
reuse, which is why day by day, they require new strategies to treat and recycle the
wastewater they generate in the different production processes [1]. Water quality is
affected by various chemical substances that dissolve in water used in each stage of the
manufacturing process.
In general, the main pollutants that are identified in the industrial wastewater are
suspended particles, organic matter, drugs, dyes, heavy metals, agrochemicals, the
hardness of the water and fats and oils [2].
The levels of concentration in which these contaminants or undesirable substances
are present in the wastewater are directly related to the operating conditions of the
productive processes. The presence of these pollutants in the water causes an impact on
the efficiency of the production processes, limits the reuse of water, increases the
consumption of clean water, the discharge of the wastewater generated contaminates the
water bodies, and this implies sanctions to the industry for exceeding the maximum
permissible limits at the effluent discharge point [3].
Currently, there are different technologies to perform the treatment of wastewater,
including adsorption using various adsorbent materials (polymers, carbon, nanomaterials,
clays, zeolites), electrodeposition, membrane filtration (ultrafiltration, nanofiltration and
reverse osmosis), coagulation-flocculation, chemical precipitation with hydroxides,
sulfides and chelating precipitations, advanced oxidation processes, electrocoagulation,
electrodialysis, ion exchange, biological treatment, photocatalysis, and electroflotation.
Each technology has certain advantages and disadvantages. Considering the most
demanding environmental legislation, industries need more efficient wastewater
treatments to eliminate suspended or dissolved metals [5-20].
In the case of coagulation-flocculation process, a very simplest equipment is required
Evaluating New Biopolyelectrolytes … 171
thus remains among the most popular process for the separation of various
types of contaminants, in which chemical substances are used as synthetic coagulant-
flocculating agents, for example, alum, ferric chloride, polyaluminum
chloride and synthetic organic polymer. The synthetic polyelectrolytes (PE) have also
been reported to pose a number of environmental problems since some of the
derivatives and intermediates are non-biodegradable and reported as hazardous to human
health (their monomers are neurotoxins and carcinogens). One of the promising
alternatives is the use of biopolyelectrolytes (BPE) as flocculating coagulants,
which can be extracted from various sources such as residues from the food industry and
agriculture.
One of the most appreciated characteristics of PE is its efficiency and effectiveness in
the destabilization of colloidal systems or dispersion of particles. In a colloidal system the
suspended particles have a surface charge density, which causes that by
electrostatic interactions the particles repel, preventing their sedimentation by gravity
forces.
Such is the case of the wastewater generated by the factory for cutting and packing of
meat products, which consists of a dispersion of particles of fats and oils from animal
origin. In this chapter, the physico-chemical characterization of meat processing
wastewater is presented. The coagulation-flocculation performance in the wastewater
treatment was evaluated using the two natural biopolyelectrolyte extracted from the
shrimp husks (chitosan) and mesquite gum. The zeta potential was used as a key
electrochemical tool to strategically dose the BPE allowing the destabilization of the
contaminants present in the wastewater.
2. EXPERIMENTAL
2.2. Zeta Potential f (pH) Profiles of the Industrial Wastewater and BPE
suspended particles and the biopolyelectrolytes, as well as the optimum dose to perform
the solid-liquid separation of the suspended particles.
Temperature (oC) 28
Alkalinity (mg/L CaCO3) 98
Fats and oils (mg/L) 15,732
TSS (mg/L) 20,000
TDS (mg/L) 3,500
SS (mL/L) 900
Turbidity (FAU) 750
EC (mS/cm) 2.00
(mV) -20
Particle size (nm) 750
Color (Pt-Co) 35,083
pH 6.8
COD (mg O2/L) 60,587
TOC (mg C/L) 53,45
BOD5 (mg O2/L) 40,250
Biodegradability (BOD5/COD) 0.67
Figure 1 the = f (pH) plot shows the charge density variation the BPE of maiz gum
(GMZ), chitosan (Ch), mesquite gum (GMT), guar gum (GG), Tule lignin (LT) and
Coffe lignin (LC) with respect to pH. The change in pH had a distinct effect with each
BPE, because of the difference between the functional groups present in the chains of the
iomacromolecules. Considering the zeta potential value (= -20 mV) of the wastewater
at pH 5.2, the biopolyelectrolyte of chitosan and mesquite gum having positive surface
charge density, were selected to carry out the strategic dosing in the wastewater with a
high content of fats and oils and to evaluate their performance in coagulation-
flocculation.
Figure 2 shows the variation of zeta potential with respect to the pH of the meat
processing wastewater and chitosan.
174 E. A. López-Maldonado and M. T. Oropeza-Guzmán
30
20
10 BPE-GMZ
0
BPE-Ch
-10
-20 BPE-GMT
(mV)
BPE-GG
-30
-40
-50 BPE-LT
-60
-70
BPE-LC
-80
-90
2 3 4 5 6 7 8 9 10 11 12
pH
Figure 1. = f (pH) profiles of GMZ, Ch, GMT, GG, LT and LC.
The zeta potential value of the wastewater shows that the suspended particles
have a negative surface charge density at pH 4-12, while at pH <4 zeta potential
values close to neutrality or positive. The = f (pH) profile show that at
pH = 3.0 the wastewater has their isoelectric point ( = 0). The stability of the
particles to remain suspended is due to the value of the negative zeta potential ( = -
25.5mV).
In order to destabilize the dispersion of fats and oil, the addition of a cationic
coagulant-flocculant agent that allows the neutralization of the negative surface charge is
necessary.
The wastewater at pH 5.4 has a = -25.5 mV and the chitosan = 15.0mV, the
dosage of chitosan at this pH by pure electrostatic interaction ensures its reaction.
As expected, in order to carry out a coagulation-flocculation process in an
efficient way, it is necessary to add a polyelectrolyte with a positive charge to neutralize
the negative charge of meat processing wastewater. The monitoring of the zeta
potential value with respect to the polyelectrolyte dose in the wastewater
allows to construct the coagulation-flocculation operation curves, this ensures
that the operators of the wastewater treatment plants avoid the problems of
overdosing of agents coagulants-flocculants and save on the consumption of chemical
substances [22].
In this chapter, the capacity of two biopolyelectrolytes for the clarification of
wastewater was evaluated. This with the aim of increasing the reuse potential of the
treated water and improving the efficiency of solid-liquid separation processes.
Evaluating New Biopolyelectrolytes … 175
In Figure 3, it is shown that the zeta potential value of the meat processing
wastewater as the dose of chitosan increases ( = -20.1mV to = -0.1mV), reaching the
isoelectric point at a dose of 68 mg/L chitosan. The water treated at pH 5.4 with chitosan
has a value of BOD5 = 30mg O2/L, fats and oil = 1mg/L, turbidity = 2 FAU, TSS =
2mg/L and COD = 50mg/L, the treated water with these physicochemical characteristics
can be discharged to the municipal sewer system or reused as process water.
OH OH
O
50 O O O
HO n
NH HO NH2
40 O
30 chitosan
20
10
(mV)
-10
-20
Meat processing
-30
wastewater
-40
2 4 6 8 10 12
pH
20
BPE-Ch (68 mg/L)
15 BPE-Ch
10
BPE-GMT (90 mg/L)
5
0
(mV)
-5
BPE-GMT
-10
-15
-20
-25
0 20 40 60 80 100 120
BPE dose (mg/L)
Additionally, Figure 3 shows the behavior of zeta potential with respect to the dose of
mesquite gum, indicating that a higher dose of this biopolyelectrolyte is required to reach
the isoelectric point of the residual water.
It should be noted that in a higher dose than the IEP of wastewater, chitosan can
reestablish particles flocculated, unlike mesquite gum. This is attributed to the fact that
mesquite gum has a lower surface charge density [23-24].
The quality of the treated water obtained with the best dose (90 mg/L) of mesquite
gum does not comply with the maximum permissible limits of COD = 650 mg/L and
BOD5 = 1200 mg/L for discharge to the sewage system, however, it does not stop to be an
attractive alternative as a physicochemical pretreatment stage for the residual water
coupled to an advanced oxidation process that allows the complete mineralization of the
dissolved substances in the treated water.
CONCLUSION
ACKNOWLEDGMENTS
The authors thank the Autonomous University of Baja California for the financing
granted for the development of this project (UABC-PTC-668).
Evaluating New Biopolyelectrolytes … 177
CONFLICT OF INTEREST
The authors state that there is no conflict of interest.
REFERENCES
[1] Muralikrishna, I. V., Manickam, V., editors, (2017). Environmental
Management:Science and Engineering for Industry. United States: Elsevier; 295 -
336.
[2] Cahoon, L. B., Mallin, M. A., (2013). Water Quality Monitoring and
Environmental Risk Assessment in a Developing Coastal Region, Southeastern
North Carolina. In: Ahuja, S., editor. Monitoring Water Quality. Amsterdam:
Elsevier; 149 - 169.
[3] Suthar, S., Sharma, J., Chabukdhara, M., Nema, A.., (2010). Water quality
assessment of river Hindon at Ghaziabad, India: impact of industrial and urban
wastewater. Environ. Monit. Assess, 165, 103 - 112.
[4] Ali, A., Saeed, K., (2015). Phenol removal from aqueous medium using chemically
modified banana peels as low-cost adsorbent. Desalin. Water Treat., 57, 11242 -
11254.
[5] Gupta, V. K., Ali, I., (2008). Removal of endosulfan and methoxychlor from water
on carbon slurry. Environ. Sci. Technol., 42, 766 - 770.
[6] Wojnarovits, L., Foldvary, C. M., Takacs, E., (2010). Radiation-induced grafting of
cellulose for adsorption of hazardous water pollutants: a review. Radiat. Phy
Chem., 79, 848 - 862.
[7] Fox, M. A., Dulay, M. T., (1993). Heterogeneous photocatalysis. Chem. Rev., 93,
341 - 357.
[8] Bromley-Challenor, K. C. A., Knapp, J. S., Zhang, Z., Gray, N. C. C., Hetheridge,
M. J., (2000) Evans MR. Decolorization of an azo dye by unacclimated activated
sludge under anaerobic conditions. Water Res., 34, 4410 - 4418.
[9] dos Santos, A. B., Cervantes, F. J., van Lier, J. B., (2007). Review paper on current
technologies for decolorisation of textile wastewaters: perspectives for anaerobic
biotechnology. Bioresour. Technol., 98, 2369 - 2385.
[10] Shen, H., Wang, Y. T., (1994). Biological reduction of chromium by E. Coli. J.
Environ. Eng., 120, 560 - 572.
[11] Ali, I., Khan, T. A., Asim, M., (2011). Removal of arsenic from water by
electrocoagulation and electrodialysis techniques. Sepn. Purfn. Rev., 40, 25 - 42.
[12] Suzuki, Y., Maezawa, A., (2000). Uchida S. Utilization of ultrasonic energy in a
photocatalytic oxidation process for treating waste water containing surfactants.
Jpn. J. Appl. Phys., 39, 2958 - 2961.
178 E. A. López-Maldonado and M. T. Oropeza-Guzmán
Chapter 9
ABSTRACT
The brewing industry presents a great generation of effluent that can cause serious
environmental impacts when not treated in a regular way since it has a high load of
organic matter in its composition. Thus, in view of the growing emergence of breweries
in Brazil and consequent increase in effluent production, alternatives are sought for the
auxiliary treatment of this effluent using coagulants. In this way, the paper aimed to
analyze the efficiency the usage of two organic and one inorganic coagulant in the
treatment of effluent from the brewing industry. The assay was performed in Jar Test,
thus simulating the coagulation, flocculation and sedimentation processes, and in
sequence the filtration process was carried out. During this process the parameters of
apparent color, electrical conductivity, pH and turbidity were monitored. Three different
treatments were used: Aluminum Sulfate (T1), Tannin (T2) and Moringa oleifera (T3).
180 F. J. L. Janz, E. R. Pereira, T. Ribeiro et al.
The results showed values of apparent color removal and turbidity of more than 90% for
all treatments, but the treatment with Tanino presented the highest values of removal,
being 98.8% for turbidity removal and 99.6% for removal of apparent color. The pH
parameter showed a rise throughout the test for all treatments presenting final values
close to the neutral value, demonstrating compliance with CONAMA Resolution No.
430/2011. The electrical conductivity showed a decrease in all treatments especially in
the treatment where Tanino was used being 0.32μS.m-1 the lowest value obtained. The
use of the coagulants in question showed positive responses in the analyzed parameters in
the treatment of the effluent of the brewing industry, showing more emphasis for the
Tanino that obtained the best values. It’s an option sustentable and is possible to mix this
process with the comum process that there are in brewing industry.
1. INTRODUCTION
The brewing market has shown a growth in Brazil, and between 2014 and 2017 there
was a 91% increase in the number of establishments registered in the country, from 356
to 679 breweries (Marcusso; Müller, 2017).
The brewing sector in Brazil is considered a relevant sector for the country’s
economy, having produced between 2011 and 2014 a turnover of 70 billion reais a year,
generating 2.2 million jobs and producing 14 billion liters of beer per year, the sector is
responsible for 1.6% of the Brazilian domestic productof the country (CERVBRASIL,
2018).
The effluent generation in the brewery varies from 3 to 6 L of effluent per liter of
beer produced, occurring in several sectors of the industry, and the lavage sector is
responsible for the largest generation of effluent, but this effluent presents low organic
load, and the fermentation and filtration sectors generate little volume of effluent, but
with high organic load (CETESB, 2005).
The effluent generated has high organic load, suspended solids content and presence
of nutrients (phosphorus and nitrogen), which causes this effluent to be considered a high
polluting potential (CETESB, 2005).
Most of the breweries have a treatment system composed of biological and
physicochemical processes, and the physical-chemical process has the function of
reducing the organic load of the effluent that will later undergo a biological process,
being aerobic or anaerobic. The organic load of the brewery effluent is present in solid
form, and its removal can be through the physical-chemical process of coagulation
(Kochenborger, 2012).
Knowing the ability of the coagulation process to remove solid materials and
improve effluent quality, the study aimed to compare the use of organic and inorganic
coagulants in the treatment generated by the brewing industry.
Treatment of Residual Waters of the Brewery Industry ... 181
2. METHODOLOGY
For the accomplishment of the study three forms of treatments were determined,
being used three different coagulants, so defined: T1): application of the Aluminum
Sulphate; (T2): application of Tannin; and (T3): application of Moringa oleifera, using
the same concentration of solution corresponding to 9mg.L-1 in all treatments.
The preparation of the Moringa oleifera solution was carried out by processing 50g
of peeled seed and 1M NaCl in 1L of distilled water. The Tannin solution was prepared
with homogenization of 1mL of coagulant in 1L of distilled water. For the preparation of
the Aluminum Sulphate solution, 1g of the chemical was homogenized in 1L of distilled
water.
The laboratory test was performed on Jar-Test equipment in order to reproduce the
coagulation, flocculation and sedimentation processes, followed by the filtration process
of the effluent. The equipment was programmed according to Theodoro (2012), initiating
the rotation of the blades at 150rpm for a time of 3 minutes, then taking the reduction to
15rpm for a period of 10 minutes and then ceasing the stirring process. The filtration
process was performed using a downflow sand filter and the filter construction was
performed according to Di Bernardo et al. (2003). The filter bed was built in the lower
part with sand with granulometry less than 0.425 mm and in the upper part with sand
with granulometry in the range of 0.425 to 0.800 mm.
The sample collections during the test were programmed to be performed every 10
minutes during the sedimentation process, and a final sample collection was performed
after the filtration process. The samples were submitted to analysis of the parameters of
apparent color, turbidity, pH and electrical conductivity according APHA (2012).
The laboratory test was performed in duplicate to avoid possible errors and the data
were analyzed statistically through the application of analysis of variance (ANOVA) and
Tukey test with a significance level of 5% with the use of BioEstat 5.0 software.
3. RESULTS
electrical conductivity value. The reduction presented by all treatments is due to the
adsorption process of the salts in the filter media used.
The electrical conductivity data were submitted to the ANOVA variance test and the
results obtained are presented in Table 1.
Statistical analysis showed that the value of P is 0.1629, the fact that P is greater than
0.05, shows that the treatments did not present significant variances with each other when
the values obtained after the filtration process were analyzed.
The pH values of all treatments remained between 4.5 and 5 before the filtration
process (Figure 2). Comparing with the value of 4.6 of the crude effluent it is possible to
observe that the coagulants did not influence the variation of this parameter. The values
collected in the samples after the filtration presented an increase of the parameter, where
T1 presented the highest increase in pH, obtaining 6.6.
All treatments presented pH values within the limit established by Resolution No.
430 of 2011 of the National Environmental Council (Brazil, 2011), which determines that
effluent must have pH between 5 and 9 for launch in any water body. The results of the
statistical analysis showed that in the ANOVA test (Table 2) that the treatments presented
differences between them, and after the Tukey test (Table 3) it was evidenced that the
difference occurs in the comparison of T1 and T2.
Variation Source GL SQ MQ F P
Between groups 2 0.058 0.029 3.5358 0.1629
Within the groups 3 0.024 0.008
Total 5 0.082
Treatment of Residual Waters of the Brewery Industry ... 183
Variation Source GL SQ MQ F P
Between groups 2 0.395 0.198 11.2273 0.04
Within the groups 3 0.053 0.018
Total 5
When analyzing the turbidity removal efficiency of all treatments before the filtration
process (Figure 3), it is possible to observe that the efficiency of the coagulants used was
related to time, that is, the efficiency increases over time, T3 was the one that presented
the best behavior, but T3 in its first instants after the application showed
increased turbidity, this occurs due to the solution being opaque and the presence of
organic matter.
After the filtration process, better values of turbidity removal efficiency were
obtained for all treatments, which were higher than 90%, with T2 (Tannin) showing the
best turbidity removal efficiency with a result of 98.8%, while T1 and T3 obtained results
of 93.5% and 94.1%, respectively. Statistical results (Table 4 and Table 5) demonstrated
that T2 presented a significant difference with the other treatments, thus highlighting its
positive performance of removal.
Variation Source GL SQ MQ F P
Between groups 2 33.699 16.849 19.9381 0.0178
Within the groups 3 2.535 0.845
Total 5
For the apparent color parameter, it can be seen from Figure 4 that
T1 and T3 treatments presented higher efficiencies than T2 treatment. With
the use of filtration, an improvement in the apparent color removal values of all
treatments was observed, thus presenting removal values above 90%. The T2 treatment
presented an increase of 18.2% (sedimentation) to 99.6% (filtration) denotand
efficiency with the filtration process, thus finalizing the test with the best result in the
parameter.
Statistical analysis showed that the procedures presented a significant difference
when ANOVA (Table 6) and Tukey’s test (Table 7) were performed. This difference was
shown in the comparison of T2 with the other treatments. The treatments T1 and T3 do
not present significant differences if.
Variation Source GL SQ MQ F P
Between groups 2 29.093 14.531 29.9624 0.0098
Within the groups 3 1.455 0.485
Total 5 30.548
CONCLUSION
The use of the three coagulants presented demonstrated efficiency for the
treatment of the effluent from the brewing industry, besides showing that the
application of organic coagulants presents final results similar or superior to the inorganic
coagulant. All treatments obtained results superior to 90% for removal of apparent color
and turbidity, but the tannin presented greater prominence presenting removal efficiency
of 98.8% of turbidity and 99.6% of apparent color. All treatments were in compliance
with CONAMA 430/2011.
REFERENCES
APHA – American Public Health Association. Standard Methods for the Examination of
Water and Wastewater. 22 ed. Washington, 2012.
186 F. J. L. Janz, E. R. Pereira, T. Ribeiro et al.
Brazil. National Council for the Environment. Resolution No. 430, of May 13, 2011.
Official Gazette of the Federative Republic of Brazil, Brasília, DF, May 16. 2011.
Available at: <http://www.mma.gov.br/port/conama/legiabre.cfm?codlegi=646>.
Accessed on: 31 Aug. 2015.
CERVBRASIL. Brazilian Association of the Beer Industry. Yearbook 2016. Available at:
http://www.cervbrasil.org.br/paginas/index.php?page=anuario-2015. Accessed on:
May 15, 2018.
CETESB. Secretariat of the Environment. Environmental Sanitation Technology
Company - CETESB. Beers and Soft Drinks. São Paulo: CETESB, 2005. Available
at: http://www.cetesb.sp.gov.br. Accessed on: May 17, 2018.
Di Bernardo, L., Mendes, C. G. N., Brandão, C. C. S., Sens, M. L., Pádua, V. L. Water
Treatment for Direct Filtration. Di Bernardo, Luiz (coordinator) - Rio de Janeiro:
ABES, RiMa, 2003. Project PROSAB 498 p.
Kochenborger, G. Physical-chemical treatment for brewery effluent. 2012. 40 f. TCC
(Undergraduate) - Course of Environmental Engineering, Passo Fundo University,
Passo Fundo, 2012.
Marcusso, E. F., Müller, C. V. Beer in Brazil: The Ministry of Agriculture Informing and
Clarifying. Ministry of Agriculture, 2017. Available at: http://www.agricultura.
gov.br/assuntos/inspecao/produtos-vegetal/pastapublicacoes-DIPOV/a-cerveja-no-
brasil-28-08.pdf. Accessed on: May 15, 2018.
Theodoro, J. D. P. Study of coagulation/flocculation mechanisms to obtain potable water
for human consumption. 2012. 185f. Doctoral Thesis (Department of Chemical
Engineering) - Technology Center, State University of Maringá, Maringá, set. 2012.
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 10
ABSTRACT
1. INTRODUCTION
This paper studied an option of water treatment, the eletrocoagulation, which consists
in the application of potential difference in an aluminum electrode, which undergoes
oxidation at the anode and reduction at the cathode, the oxygen microbubbles formed by
the cathode help that the pollutant particles are suspended (Martins 2017).
According to Adapureddy (2012), the initial pH influences the removal parameters
for water purification or effluent treated by electrocoagulation. His job shows that the
initial pH around 8 was the best results of removal parameters, and he also observed that
at pH ≥ 9 the generation of microbubbles increases which shows more gas production.
The electrical conductivity is an indicator of the amount of dissolved salts, however it
is possible to measure the presence of salts in the collected water since in this work no
salts will be added for an acceleration of the process. Therefore, the objective of this
paper is to analyze the parameters of pH and of conductivity in the water treatment
process by electrocoagulation, and also to calculate the cost of this operation.
2. METHODOLOGY
The sampling was made at the water inlet tank of the treatment station (Figure 1),
located in Paraná, Brazil.
Figure 2. Bench level reactor: A) Monopolar electrode system; B) Recipient of the reactor; C) Magnetic
stirrer; and D) Continuous voltage source.
Initial pH
Tests Time (min)
5V 11V
1 40 4.06 4.03
2 60 4.06 4.03
3 80 4.06 4.03
4 40 6.31 6.31
5 60 6.31 6.31
6 80 6.31 6.31
7 40 8.80 9.07
8 60 8.80 9.07
9 80 8.80 9.07
10 60 6.31 6.31
11 40 4.06 4.03
12 60 4.06 4.03
13 80 4.06 4.03
14 40 6.31 6.31
15 60 6.31 6.31
16 80 6.31 6.31
17 40 8.80 9.07
18 60 8.80 9.07
19 80 8.80 9.07
20 60 6.31 6.31
190 I. C. Botelho, Th. A. Bezerra Higuchi and J. D. Peruço Theodoro
Table 2. pH corretion
For the analysis of pH and conductivity, were collected 25mL samples in 40,
60 and 80 minutes; the reading was made straight after collecting them. In the
statistical planning (Table 1) it was performed 9 trials in duplicate and a pair in the
central point.
The methodology and equipment used in the parameter readings pH and
conductivity were, respectively, Potentiometric Method range 1.0 to 13.0 and
pHmetro mPA-210 and LQ Electrometric Method: 0.01μS.cm-1 and Mca 150
Conductor (Standard Methods for the Examination of Water and Wastewater (American
Public Health Association, 2012). The acid and basic pH were corrected using solutions
of hydrochloric acid (1M HCl) and sodium hydroxide (1M NaOH), respectively (Table
2).
In relation to the monopolar electrode system, these were connected in parallel and
fixed by steel screws, the four aluminum plates have 2cm equidistance and dimensions
16cm high x 8cm long x 0.03cm thick, resulting in two pairs of electrodes. The adjustable
voltage source allowed the current to pass through the system and the magnetic stirrer
was used to homogenize the fluid.
On the other hand the use of the electrode (1) and the energy consumption (2) were
calculated by de equations below.
i.t.M
𝑚𝑒 = n.F
(1)
where:
𝑚𝑒 Is an amount of dissolved electrode (g);
Application of Electrocoagulation with Voltage Variation ... 191
i Is current (A);
t The process time (s);
M The molar ratio of the predominant element of the electrode (g.mol-1);
n Is the number of electrons in the oxidation reaction of the anodic electrode; and F It
is a Faraday constant (F = 9.65 x 104 C.mol-1).
𝑈.𝑖.𝑡
𝐶= (2)
𝑣
where:
𝐶 Energy consumption (Wh.m-³);
𝑈 Applied electric voltage (V);
𝑖 Electric current (A);
𝑡 Time of application of the current or process time (h); and
𝑣 Volume of treated effluent (m³)
Through the results found with the two equations above, the cost of the process was
calculated using equation (3).
where:
𝐶𝑜𝑝 Total system operating cost (R$.m-3);
𝑎 Cost of electricity (R$.kWh-1);
𝐶𝑒𝑛𝑒𝑟𝑔 Consumption of electric energy (kWh.m-3);
𝑏 Cost of the electrodes (R $.kg aluminum-1); and
𝐶𝑒𝑙𝑒𝑡 Electrode consumption (kg.m-3 water treated).
3. RESULTS
The characterization values of the crude sample are shown in Table 3. In addition, the
results of the experimental design for the electrocoagulation process are
presented in Table 4.
Parameters Values
pH 5.99
Conductivity (μS.cm-1) 0.17
192 I. C. Botelho, Th. A. Bezerra Higuchi and J. D. Peruço Theodoro
Conductivity
Initial pH Final pH
Tests Time (min) (μS.cm-1)
5V 11V 5V 11V 5V 11V
1 40 4.06 4.03 5.47 4.95 0.18 0.17
2 60 4.06 4.03 5.74 5.1 0.18 0.17
3 80 4.06 4.03 5.52 5.47 0.18 0.17
4 40 6.31 6.31 5.59 6.52 0.16 0.16
5 60 6.31 6.31 5.66 6.41 0.18 0.16
6 80 6.31 6.31 5.88 6.87 0.16 0.15
7 40 8.80 9.07 7.25 7.74 0.18 0.20
8 60 8.80 9.07 7.38 7.32 0.18 0.20
9 80 8.80 9.07 7.38 7.13 0.18 0.21
10 60 6.31 6.31 5.79 6.22 0.16 0.17
11 40 4.06 4.03 4.94 5.54 0.20 0.18
12 60 4.06 4.03 4.77 5.16 0.20 0.18
13 80 4.06 4.03 4.87 5.11 0.20 0.17
14 40 6.31 6.31 5.61 6.25 0.17 0.16
15 60 6.31 6.31 6.05 6.63 0.17 0.15
16 80 6.31 6.31 6.02 6.74 0.16 0.15
17 40 8.80 9.07 6.99 7.58 0.18 0.19
18 60 8.80 9.07 6.73 7.45 0.18 0.18
19 80 8.80 9.07 6.64 7.61 0.18 0.18
20 60 6.31 6.31 5.88 6.44 0.18 0.16
Figure 3 shows the graphic of final pH variation, the analysis of the results
confirmed what Adapureddy says, that the electrocoagulation process brings
the water near to neutral pH. At the conventional water treatment, after the
addition of the coagulant the pH decrease, thus the neutralization of water is required.
194 I. C. Botelho, Th. A. Bezerra Higuchi and J. D. Peruço Theodoro
However, the study shows that the electrocoagulation process tends to keep the water pH
neutral.
The variation of electrical conductivity represented in Figure 4 had few variations, it
can be interpreted by no salts addition moreover there aren’t many changes at the
conductivity by variation of 5V and 11V voltage.
The last analysis was to determine the cost of the operation, the beginning of the
electrode use was calculated (Table 5), then the energy consumption (Table 6) and then
with these values and the prices of the material and kWh, the total cost was found
(Table 7).
CONCLUSION
In the analysis of pH and the electrical conductivity behavior, it was observed that the
electrocoagulation assists in the neutrality of the solution, and for electrical conductivity
the process does not show significant changes, maintaining the range of 0.15 to
0.20μS.cm-1.
The operation cost in the applied voltage of 5V is lower, therefore more interesting to
use this voltage to the eletrocoagulation process, which consumes less energy, making the
experiment more satisfactory.
REFERENCES
Chapter 11
ABSTRACT
*
Corresponding Author Email: [email protected].
198 Ayşe Büşra Şengül
sulphate, ferric chloride, and ferrous sulphate, respectively. On the other hand, for all
coagulants, the optimum polyelectrolyte dosage for the removal of both microcystins and
turbidity was found to be 0.2 mg/L. Among these coagulants, aluminum sulphate (65.2%)
was found to be more effective for both microcystins and turbidity removal compared to
ferric chloride (56.3%) and ferrous sulphate (30.9%). Last, the concentration of turbidity
decreased to below 5 NTU, which is the maximum turbidity level in drinking water
determined by the World Health Organization (WHO). After coagulation/flocculation,
the total microcystin concentration (4.09 µg/L) in the water remained above the WHO
guideline value of 1 µg/L for drinking water, but the release of toxins in the water was
not observed. Therefore, both intracellular microcystin and turbidity could be removed
through the coagulation/flocculation process, whereas extracellular microcystin is still of
particular concern in drinking water treatment due to its difficult removal.
INTRODUCTION
Organization (WHO) has specified a guideline value for microcystins in drinking water
as 1 µg/L (WHO, 2003).
Generally, microcystins are contained within cells that are classified as cell-bound or
intracellular (Campinas & Rosa, 2010b, c; Dixon et al., 2011). When the cells are
subjected to natural toxin release (natural cell lysis and death, active release) or induced
toxin release (mechanical/chemical stresses), microcystins can be released into
surrounding waters, and these are classified as dissolved or extracellular toxins (Schmidt
et al., 2002; Campinas & Rosa, 2010b; Dixon et al., 2011; Li et al., 2015). Due to the fact
that extracellular microcystins are not removed as easily as intracellular toxins (Sun et al.,
2012; Sun et al., 2013), it is necessary to develop an efficient method to remove
intracellular microcystins without cell damage.
Coagulation/flocculation is one of the most common processes to improve treatment
efficiency and cost effectiveness for the removal of cyanobacterial cells and intracellular
toxins (Pei et al., 2014; Jiang, 2015). A number of researchers have been focusing on the
impacts of the coagulation/flocculation process on cyanobacterial cells and toxin release
(Han et al., 2012; Sun et al., 2012; Han et al., 2013; Sun et al., 2013; Pei et al., 2014; Li
et al., 2015; Ma et al., 2016). However, in the literature, the coagulation/flocculation
process for the uptake of cyanobacterial cells and toxin is shown to be controversial.
Therefore, some research has been reported that the optimized coagulant dosage as well
as mechanical conditions did not damage the membrane of cyanobacterial cells and the
release of cell metabolites (Chow et al., 1998; Chow et al., 1999; Sun et al., 2012; Sun et
al., 2013; Pei et al., 2014; Li et al., 2015). Other research has emphasized that chemical
treatment with maximum dosage and long contact time causes cell damage, thus releasing
large amounts of toxin into the water (Han et al., 2012; Han et al., 2013). These findings
clearly show that the efficiency of the coagulation/flocculation process can be strongly
affected by the type and dosage of coagulant, pH values, applied shear, and
characteristics of the raw water.
Up to now, a limited number of studies have been assessing the effects of the
coagulation/flocculation process for these studies have only focused on the impact of the
coagulant dosage and contact time for the toxin removal. The literature survey has
demonstrated that there has not been a study on intracellular and extracellular microcystin
removal relative to the influence of pH and coagulant-aid dosage in addition to a
coagulant. Therefore, the main objective of this study was to examine the performance of
the coagulation/flocculation process using aluminum sulphate, ferric chloride, and ferrous
sulphate coagulants for the removal of both microcystins (intracellular and extracellular)
and turbidity. The effects of coagulant dosage, pH, and coagulant-aid dosage on
intracellular and extracellular microcystins and turbidity removal were also thoroughly
investigated.
200 Ayşe Büşra Şengül
Algae Culture
Natural Water
Natural water was collected from Büyükçekmece Lake (a drinking water source in
Istanbul, Turkey). After filtering through a 0.45 µm filter, cultures in the BG-11 medium
were transferred to the natural water. Characteristics of this natural water are provided in
Table 1.
Chemicals
mg/L) to determine the optimum coagulation conditions. The pH was varied between 5
and 8 by adding 0.1 M sodium hydroxide (NaOH) and 0.1 M hydrochloric acid (HCl). A
coagulant aid, cationic polyelectrolyte, was used as a 1g/L stock solution and added to the
beaker in a range of dosages (0.2, 0.4, 0.6, 0.8, and 1 mg/L).
Coagulation/Flocculation Experiment
Analytical Method
Turbidity
After settling, turbidity of the supernatant was measured using a turbidimeter (WTW
TURB 550).
Figure 1 shows the results of microcystins and turbidity removal after the
coagulation/flocculation process under various ferric chloride dosages, pH values, and
polyelectrolyte dosages. The removal of intracellular microcystin increased by increasing
the ferric chloride dosage from 0 to 60 mg/L. At a dosage of more than 60 mg/L,
intracellular microcystin concentration slightly decreased, and no further significant
removal was achieved. The concentration of intracellular microcystin decreased from 10
to 5.11 µg/L, and its removal efficiency was 48.9% in the presence of 60 mg/L ferric
chloride. The trend of turbidity removal by coagulation/flocculation was similar to that of
intracellular microcystin removal. Turbidity concentration decreased from 9 to 0.93
NTU, with 89.70% removal efficiency when the ferric chloride dosage was 60 mg/L. On
the other hand, no significant change in the amount of extracellular microcystin was
observed. Therefore, the optimum ferric chloride dosage was determined to be 60 mg/L
for both microcystins and turbidity removal.
As shown in Figure 1(b), removal of the intracellular microcystin increased by
increasing the pH from 5 to 6, but there was no significant change in its removal when
the pH level increased from 7 to 8. Turbidity removal showed a similar trend with
intracellular microcystin. Thus, the high performance of ferric chloride for removing
intracellular microcystin and turbidity was determined to be at pH 6. The concentration of
intracellular microcystin was decreased from 10 to 4.65 µg/L, with 53.5% removal
efficiency in the presence of a 60 mg/L ferric chloride dosage at pH 6. Turbidity
concentration decreased from 9 to 1.21 NTU, with a removal efficiency of 86.6%. In
An Evaluation of the Performance of the Coagulation/Flocculation … 203
addition, the removal of extracellular microcystin was not impacted when the pH level
was changed. The removal efficiency of intracellular microcystin by ferric chloride
coagulation was increased 8.6% with a change in the pH, whereas the removal efficiency
of turbidity was decreased.
The effect of polyelectrolyte dosage on microcystins and turbidity removal was
evaluated with a 60 mg/L ferric chloride dosage and pH level 6 under various dosages of
polyelectrolytes, as shown in Figure 1(c). The removal of intracellular microcystin and
turbidity increased by increasing the polyelectrolyte dosage from 0 to 0.2 mg/L; however,
no remarkable increase was obtained with a further increment in the polyelectrolyte
dosage beyond 0.2 mg/L. The concentration of intracellular microcystin decreased from
10 to 4.37 µg/L, with 56.30% removal efficiency. Turbidity concentration decreased from
9 to 2.07 NTU, with removal efficiency of 77%. Moreover, no significant changes were
observed in the amount of extracellular microcystin concentration. The removal
efficiency of intracellular microcystin by ferric chloride coagulation was increased 13.1%
with a change in the pH and the addition of a polyelectrolyte, whereas the removal
efficiency of turbidity was decreased.
Until now, only two studies in the literature have evaluated the effect of ferric
chloride coagulation/flocculation on Microcystis aeruginosa cells and
toxin release (Chow et al., 1998; Li et al., 2015), which reported that
coagulation/flocculation with ferric chloride achieved up to 100% removal of
intracellular microcystin. Also, ferric chloride and mechanical action did not damage
cyanobacterial cells and the release of extracellular toxin. In this study, only 56.3%
intracellular microcystin removal was achieved, and no further increase in the amount of
extracellular microcystin was observed, with its concentration measured at 0.61 µg/L
over the experimental period. In general, the predominant removal
mechanism at relatively low doses and slightly acidic pH value of coagulants is charge
neutralization.
However, at higher dosages and approximately neutral-to-alkaline pH values sweep
flocculation occurs (Duan & Gregory, 2003). Therefore, all results clearly showed that
the iron ions can be transformed to Fe(OH)3 by hydrolysis at pH 6 with a dosage of 60
mg/L, and the Fe(OH)3 tends to remove intracellular microcystin via sweep flocculation.
Nevertheless, natural organic matter in aquatic systems may result in lower
intracellular microcystin removal efficiency (Ma & Liu, 2002) and extracellular
microcystin are efficiently removed by charge neutralization more than sweep
coagulation (Ghernaout, 2014). Another reason is that when the initial concentration of
microcystin is around 10 µg/L or higher, there may not be a significant reduction in the
microcystin concentration.
204 Ayşe Büşra Şengül
Figure 2 demonstrates the jar test results of microcystins and turbidity removal after
the coagulation/flocculation process under various ferrous sulphate dosages, pH values,
and polyelectrolyte dosages. The removal of intracellular microcystin increased with a 20
mg/L dosage of ferrous sulphate. Then, the removal of intracellular microcystin gradually
decreased with the continuing increase in ferrous sulphate dosages. The concentration of
intracellular microcystin decreased from 10 to 6.98 µg/L, and its removal efficiency was
An Evaluation of the Performance of the Coagulation/Flocculation … 205
30.2% in the presence of a 20 mg/L ferrous sulphate dosage. A similar trend was also
observed using ferrous sulphate at a dosage of 20 mg/L. Turbidity concentration
decreased from 9 to 2.53 NTU, with a removal efficiency of 71.9%. No significant
turbidity occurred, after a dosage of 20 mg/l. On the other hand, the amount of
extracellular microcystin was not changing. Hence, 20 mg/L ferrous sulphate dosage was
considered as optimum dosage due to the highest obtained microcystin and turbidity
removal.
As shown in Figure 2(b), removal of the intracellular microcystin was increased by
increasing the pH from 5 to 7, but no significant change was observed for both
microcystin and turbidity removal when the pH level increased from 7.5 to 8. Turbidity
removal showed a similar trend with intracellular microcystin. Thus, the high
performance of ferrous sulphate in removing intracellular microcystin and turbidity was
determined to be at pH 7.
The concentration of intracellular microcystin was decreased from 10 to 5.43
µg/L, with 45.7% removal efficiency in the presence of a 20 mg/L ferrous sulphate
dosage at pH 7. Turbidity concentration decreased from 9 to 3.26 NTU, with a
removal efficiency of 63.78%. In addition, the removal of extracellular microcystin was
not clearly impacted by changing the pH level. The efficiency of removing
intracellular microcystin using ferrous sulphate coagulation was increased
33.9% under different pH levels, unlike the efficiency of removing turbidity, which
decreased.
The effect of the polyelectrolyte dosage on the removal of microcystins and turbidity
was evaluated with an optimized ferrous sulphate dosage and pH under
various polyelectrolyte dosages, as shown in Figure 2(c). The removal of intracellular
microcystin and turbidity increased by increasing the polyelectrolyte dosage
from 0 to 0.2 mg/L. However, results showed that when the polyelectrolyte
dosage was further increased from 0.2 to 1 mg/L, no additional obvious increase was
observed.
The concentration of intracellular microcystin decreased from 10 to 6.91 µg/L, with
30.9% removal efficiency. The turbidity concentration decreased from 9 to 3.62 NTU,
with 59.8% removal efficiency. Moreover, no significant increase in the amount of
extracellular microcystin concentration was observed with the addition of
polyelectrolytes. The removal efficiency of intracellular microcystin and turbidity by
ferrous sulphate coagulation was decreased with a change in the pH and the addition of
polyelectrolytes.
In this study, ferrous sulphate showed poor intracellular microcystin and turbidity
removal efficiency compared to aluminum sulphate and ferric chloride. The maximum
intracellular microcystin removal for ferrous sulphate was only 45.7% in the presence of
a 20 mg/L ferrous sulphate dosage at pH 7. When a higher dosage of ferrous sulphate was
206 Ayşe Büşra Şengül
added, the effectiveness of the processes decreased further. There was no change in the
removal of extracellular microcystin.
(a)
(b)
(c)
Jar test experiments were performed with aluminum sulphate, ferric chloride, and
ferrous sulphate coagulants to evaluate the performance of the coagulation/flocculation
process based on microcystins and turbidity removal (Figure 3). The concentration of
intracellular microcystin decreased from 10 µg/L to 3.48, 4.37, and 6.91 µg/L, with 65.2,
56.3, and 30.9% removal efficiency for aluminum sulphate, ferric chloride, and ferrous
sulphate coagulants, respectively. The optimum dosage of aluminum sulphate, ferric
chloride, and ferrous sulphate for maximum microcystins and turbidity removal was
found to be 60, 60, and 20 mg/L respectively. The corresponding turbidity removal
An Evaluation of the Performance of the Coagulation/Flocculation … 207
efficiencies of aluminum sulphate, ferric chloride, and ferrous sulphate were found to be
85.1, 77.7, and 59.8%, respectively. The concentration of turbidity decreased from 9
NTU to 1.34, 2.07, and 3.62 NTU for aluminum sulphate, ferric chloride, and ferrous
sulphate, respectively.
AS FC FS
AS_T FC_T FS_T
Turbidity (NTU)
Microcystins %
100 10
80 8
Removal
60 6
40 4
20 2
0 0
20 40 60 80 100
Coagulant Dosage (mg/L)
(a)
AS FC FS
AS_T FC_T FS_T
Microcystins %
Turbidity (NTU)
100 10
Removal
80 8
60 6
40 4
20 2
0 0
5 6 7 7.5 8
pH
(b)
AS FC FS
AS_T FC_T FS_T
Turbidity (NTU)
Microcystins %
100 10
80 8
Removal
60 6
40 4
20 2
0 0
0.2 0.4 0.6 0.8 1
Figure 3. Removal of microcystins (intracellular and extracellular) and turbidity removal after
coagulation/flocculation process under different conditions: (a) coagulant dosage, (b) pH value, and (c)
polyelectrolyte dosage.
An excessive dosage of coagulant caused the reversal of surface charge and the
restabilization of particles, which led to a decrease in the microcystins and turbidity
208 Ayşe Büşra Şengül
removal efficiency. With the optimized coagulant dosages, the pH for maximum
microcystins removal occurred at a pH of 7, 6, and 7 for aluminum sulphate, ferric
chloride, and ferrous sulphate, respectively, while the pH change with optimized
coagulant dosage led to a decrease in turbidity removal efficiency. Furthermore, the
polyelectrolyte dose increased the removal efficiency of microcystins, whereas it
decreased the removal efficiency of turbidity.
Overall, results indicated that aluminum sulphate was found to be a more effective
coagulant for both microcystin and turbidity removal compared to ferric chloride and
ferrous sulphate coagulants. Also, ferrous sulphate was found to be the least effective of
the other coagulants. Turbidity concentration decreased below 5 NTU, which is the
maximum level of turbidity in drinking water determined by the WHO after the
coagulation/flocculation process with all coagulants. However, microcystin concentration
in the water did not achieve the WHO guideline value of below 1 µg/L for drinking
water. Coagulation/flocculation processes achieved limited ability to remove the
microcystins, possibly because of the natural organic matter and the initial concentration
of microcystins in the water.
LITERATURE REVIEW
Coagulation/flocculation is still one of the essential treatment processes for algae and
cyanobacteria, and the associated removal of their metabolites (Ma & Liu, 2002;
Gonzalez-Torres et al., 2014). This process involves the addition and mixing of
coagulants to destabilize negative charges of particles and to prevent electrostatic
repulsion between particles. Therefore, destabilized particles can tend to agglomerate and
form flocs that are subsequently removed by sedimentation (Merel et al., 2013; Sun et al.,
2013). Coagulation/flocculation in water treatment occurs by several mechanisms: charge
neutralization, adsorption and bridging between particles, electrostatic patch, and sweep
flocculation (Duan & Gregory, 2003; Ghernaout & Ghernaout, 2012; Gonzalez-Torres et
al., 2014). Intracellular toxins can be removed by coagulation, via either charge
neutralization (requires acidic pH values and lower dosages) or sweep flocculation
(requires neutral to alkaline pH values and high dosages), and subsequent
sedimentation and filtration processes (Jiang & Kim, 2008; Wu et al., 2011).
Extracellular toxins are removed more efficiently by charge neutralization than
by sweep flocculation (Ghernaout, 2014). However, the efficiency of intracellular or
extracellular toxin removal through coagulation can be strongly affected by the type and
dosages of coagulant, applied shear, solution pH, and characteristics of water,
particularly natural organic matter in water (Hitzfeld et al., 2000; Ma & Liu, 2002;
Zhang et al., 2014).
Table 2. Summary of studies investigating performance of coagulation/flocculation process
on cyanobacterial cells and toxin release
Percent Removal
Cyanobacteria Coagulant Mechanical Action Comments References
Intracellular Extracellular
Microcystis CTS, AC, Rapid mix: 97.3 53.08 Best MC removal obtained when CTSAC set to 2.6 (Ma et al.,
aeruginosa and CTSAC 200 rpm 2 min mg/L CTS plus 7.5 mg/L AC. 2016)
Slow mix: Negligible cell lysis observed during CTS, AC, and
20 rpm 15 min CTSAC coagulation processes.
Settling time:
30 min
Microcystis Ferric Rapid mix: > 80 38.5 Optimum coagulant dosage determined as 50 mg/L. (Li et al.,
aeruginosa Chloride 250 rpm 0.5 min Coagulant dosage and mechanical actions did not 2015)
Slow mix: cause any damage to cell integrity.
20 rpm 30 min FeCl coagulant had no effect on extracellular MCs
Settling time: with coagulant dosage below 50 mg/L, but obvious
30 min removal observed at coagulant dosage of 100 mg/L.
During flocs storage process, number of intracellular
MCs released into supernatant, but cells remained
viable up to 10 d.
Microcystis CTS Rapid mix: 99.1 46.45 Optimum extracellular MCs removal efficiency (Pei et al.,
aeruginosa 226.64 rpm 1.098 obtained with chitosan concentration of 7.31 mg/L. 2014)
min Chitosan flocculation showed MCs adsorption ability.
Slow mix: During flocs storage period, MCs release observed
18.86 rpm 11.73 after 4 d.
min
Settling time:
30 min
Microcystis PACI Rapid mix: 94.3 ~ 12.0 Optimum coagulant dosage found to be 4 mg/L. (Sun et al.,
aeruginosa 150 rpm 2 min During floc storage process (after 2 days), PACl caused 2013)
Slow mix: obvious damage to cells and led to large amount of MC
40 rpm 30 min release.
Settling time:
20 min
Microcystis AS – 95 – Alum treatment with maximum dosage (48 mg/L) (Han et al.,
ichthyoblabe and long contact time (7 days) caused Microcystis 2013)
cell damage, resulting in large amounts of toxic
microcystin released into water.
Table 2. (Continued)
Percent Removal
Cyanobacteria Coagulant Mechanical Action Comments References
Intracellular Extracellular
Microcystis PAC, Rapid mix: > 99 Coagulation treatment removed Microcystis (Chen
aeruginosa Sepiolite and 200 rpm 1 min aeruginosa cells efficiently, and combined use of et al., 2013)
PAC/ Slow mix: PAC and sepiolite had higher removal efficiency
Sepiolite 50 rpm 10 min compared to each coagulant alone.
Settling time: Coagulation treatment with PAC (250 mg/L) +
– sepiolite (2.8 mg/L) or sepiolite (7 mg/L) had no
negative effect on Microcystis aeruginosa cells;
however, with PAC treatment, low level of MC
detected in water.
Microcystis AC Rapid mix: > 99 – All cells removed without damage to membrane (Sun et al.,
aeruginosa 250 rpm 1 min integrity under optimum coagulant dosage (15 mg/L 2012)
Slow mix: AlCl3).
20 rpm 20 min When flocs were stacked over 6 days, cells lysed and
Settling time: microcystin-LR concentration increased (nearly
30 min 41.7%).
Microcystis AS Rapid mix: – – Small dosage (7 mg/L as Al) of alum treatment not (Han et al.,
ichthyoblabe 300 rpm 1 min effective in precipitating Microcystis cells, and large 2012)
Slow mix: dosage (14 mg/L as Al) of alum treatment caused cell
50 rpm 5 min damage and released toxins over 80% of initial
Settling time: intracellular.
–
Anabaena AS Rapid mix: 88 – No significant turbidity occurred after dosage of 65 (Hoko &
circinalis and 200 rpm 1 min mg/l, and no further significant algae removal Makado,
Microcystis Slow mix: achieved after dosage of 90 mg/l. 2011)
aeruginosa 20 rpm 15 min Algae removal by coagulation found to be higher at
Settling time: lower pH (pH 7).
30 min
Chlorella CTS Rapid mix: 99.03 ± 0.7% – Optimal chitosan dosage for destabilizing algae cells (Ahmad
spirulina 150 rpm 2 min was 10 ppm with of 99.3 ± 0.7% removal efficiency. et al., 2011)
Settling time: Concentrations higher than 10 ppm restabilized
20 min microalgae cultures, thus reducing process efficiency.
Microcystis AS, PACl, Rapid mix: – – Dosage of 0.04 mM, PACl (90%) achieved better (Jiang &
aeruginosa Bent, Na- 400 rpm 1 min performance than AS (75%) with respect to algal Kim, 2008)
Bent, Mont- Slow mix: removal.
KSF, and 35 rpm 20 min Dosage of 200 mg/L, 100% chlorophyll-a removal
Mont-K10 Settling time: obtained by Bent and Mont-KSF clays.
60 min
Percent Removal
Cyanobacteria Coagulant Mechanical Action Comments References
Intracellular Extracellular
Microcystis AS Rapid mix: – – All cells removed without damage to membrane (Chow
aeruginosa 20 rpm 1 min integrity; thus, aluminum sulphate (5.3 mg/l as Al at et al., 1999)
Slow mix: pH 6.7) and mechanical action did not cause any
25 rpm 14 min damage to cultured M. aeruginosa cells.
Settling time:
15 min
Microcystis Ferric – – – Cultured M. aeruginosa and A. circinalis cells not (Chow
aeruginosa Chloride damaged by ferric chloride treatment; however, ferric et al., 1998)
chloride seemed to stimulate growth of both M.
aeruginosa and A. circinalis.
AC: aluminum chloride; AS: aluminum sulphate; Bent: bentonite; CTS: chitosan; CTSAC: chitosan-aluminum chloride; Mont-K10: montmorillonite; K10Mont-KSF:
montmorillonite KSF; Na-Bent: sodium modified bentonite PAC: polyaluminum chloride; PACI; polyaluminum chloride
212 Ayşe Büşra Şengül
Many studies have investigated the effect of the coagulation/ flocculation process on
cyanobacterial cells and cyanotoxin release during water treatment. Results of several
investigations into the efficiency of this process to remove cyanobacterial
cells have suggested that coagulation is more effective for the removal of
intracellular toxins than extracellular toxins, due to their hydrophilicity. Cyanobacterial
cells generally possess a negative surface charge at most pH levels, and many
common species are less than 10 mm in diameter; therefore, surface charge
can be neutralized by introducing coagulants. These coagulants can easily
flocculate algal cells and form flocs (Henderson et al., 2010; Dong et al., 2014; Fast et al.,
2014).
Nevertheless, effects of the coagulation/flocculation process on cyanobacterial
cell lysis in the literature are debatable. For example, Chow et al. (1998; 1999) evaluated
the effect of coagulation on cells of Microcystis aeruginosa and Anabaena circinalis, and
reported that chemicals and mechanical action did not damage cyanobacterial cells and
the release of cell metabolites. Sun et al. (2012) also reported that coagulant
dose and shear under optimum conditions did not cause the lysis of cells and ensuing
release of microcystin-LR.
Furthermore, Pei et al. (2014), Sun et al. (2013), and Li et al. (2015) documented that
the optimized coagulants as well as mechanical conditions did not damage the
membrane of the cultured M. aeruginosa cells. However, Han et al. (2012; 2013) found
that alum treatment with the maximum dosage and a long contact time caused
Microcystis cell damage, and as a result, large amounts of toxic microcystin were
released into the water.
Table 2 summarizes the effectiveness of different coagulation/ flocculation
conditions in removing intracellular and extracellular toxins of several of the most
important cyanobacteria.
CONCLUSION
The highest intracellular microcystin removal was achieved using the aluminum
sulphate (65.2%) coagulant in comparison with ferric chloride (56.3%) and
ferrous sulphate (30.9%) under the determined optimum conditions.
An Evaluation of the Performance of the Coagulation/Flocculation … 213
The highest turbidity removal was found to be 85.1%, 77%, and 59.8% for
aluminum sulphate, ferric chloride, and ferrous sulphate, respectively.
The pH and addition of polyelectrolyte had less effect on microcystins and
turbidity removal.
The concentration of turbidity decreased to below 5 NTU, which is the maximum
turbidity level in drinking water determined by the WHO when the
coagulation/flocculation process is applied.
The total microcystin concentration (4.09 µg/L) in the water remained above the
WHO guideline value of 1 µg/L for drinking water, but the release of toxins in
the water was not observed after the coagulation/flocculation process.
Overall, results indicate that both intracellular microcystin and turbidity could be
removed through the coagulation/flocculation process, whereas extracellular
microcystin is still of particular concern in drinking water treatment due to its
difficult removal.
REFERENCES
Ahmad, A., Yasin, N. M., Derek, C. & Lim, J. (2011). Optimization of microalgae
coagulation process using chitosan. Chemical Engineering Journal, 173(3), 879-882.
Campinas, M. & Rosa, M. J. (2010a). Assessing PAC contribution to the NOM fouling
control in PAC/UF systems. Water research, 44(5), 1636-1644.
Campinas, M. & Rosa, M. J. (2010b). Evaluation of cyanobacterial cells removal and
lysis by ultrafiltration. Separation and Purification Technology, 70(3), 345-353.
doi:10.1016/j.seppur.2009.10.021.
Campinas, M. & Rosa, M. J. (2010c). Removal of microcystins by PAC/UF. Separation
and Purification Technology, 71(1), 114-120. doi:10.1016/j.seppur.2009.11.010.
Chen, X., Xiang, H., Hu, Y., Zhang, Y., Ouyang, L. & Gao, M. (2013). Fates of
Microcystis aeruginosa cells and associated microcystins in sediment and the effect
of coagulation process on them. Toxins, 6(1), 152-167.
Chorus, I. & Bartram, J. (1999). Toxic cyanobacteria in water: A guide to their public
health consequences, monitoring and management: Spon Press.
Chow, C., House, J., Velzeboer, R., Drikas, M., Burch, M. & Steffensen, D. (1998). The
effect of ferric chloride flocculation on cyanobacterial cells. Water research, 32(3),
808-814.
Chow, C. W., Drikas, M., House, J., Burch, M. D. & Velzeboer, R. (1999). The impact of
conventional water treatment processes on cells of the cyanobacterium Microcystis
aeruginosa. Water research, 33(15), 3253-3262.
214 Ayşe Büşra Şengül
D’Anglada, L. & Strong, J. (2015). Drinking water health advisory for the
cyanobacterial microcystin toxins. Health and Ecological Criteria Division, Ed.
United States Environmental Protection Agency: Washington, DC.
Dixon, M. B., Richard, Y., Ho, L., Chow, C. W., O’Neill, B. K. & Newcombe, G. (2011).
A coagulation–powdered activated carbon–ultrafiltration–Multiple barrier approach
for removing toxins from two Australian cyanobacterial blooms. Journal of
Hazardous Materials, 186(2), 1553-1559.
Dong, C., Chen, W. & Liu, C. (2014). Flocculation of algal cells by amphoteric chitosan-
based flocculant. Bioresource technology, 170, 239-247.
Duan, J. & Gregory, J. (2003). Coagulation by hydrolysing metal salts. Advances in
colloid and interface science, 100, 475-502.
Fast, S. A., Kokabian, B. & Gude, V. G. (2014). Chitosan enhanced coagulation of algal
turbid waters–Comparison between rapid mix and ultrasound coagulation methods.
Chemical Engineering Journal, 244, 403-410.
Ghernaout, D. (2014). The hydrophilic/hydrophobic ratio vs. dissolved organics removal
by coagulation–A review. Journal of King Saud University-Science, 26(3), 169-180.
Ghernaout, D. & Ghernaout, B. (2012). Sweep flocculation as a second form of charge
neutralisation—a review. Desalination and Water Treatment, 44(1-3), 15-28.
Gonzalez-Torres, A., Putnam, J., Jefferson, B., Stuetz, R. & Henderson, R. (2014).
Examination of the physical properties of Microcystis aeruginosa flocs produced on
coagulation with metal salts. Water research, 60, 197-209.
Han, J., Jeon, B. s. & Park, H. D. (2012). Cyanobacteria cell damage and cyanotoxin
release in response to alum treatment. Water Science & Technology: Water Supply,
12(5), 549. doi:10.2166/ws.2012.029.
Han, J., Jeon, B. S., Futatsugi, N. & Park, H. D. (2013). The effect of alum coagulation
for in-lake treatment of toxic Microcystis and other cyanobacteria related organisms
in microcosm experiments. Ecotoxicol Environ Saf, 96, 17-23. doi:10.1016/
j.ecoenv.2013.06.008.
Henderson, R. K., Parsons, S. A. & Jefferson, B. (2010). The impact of differing cell and
algogenic organic matter (AOM) characteristics on the coagulation and flotation of
algae. Water research, 44(12), 3617-3624.
Hitzfeld, B. C., Höger, S. J. & Dietrich, D. R. (2000). Cyanobacterial toxins: removal
during drinking water treatment, and human risk assessment. Environmental health
perspectives, 108(Suppl 1), 113.
Hoko, Z. & Makado, P. K. (2011). Optimization of algal removal process at Morton
Jaffray water works, Harare, Zimbabwe. Physics and Chemistry of the Earth, Parts
A/B/C, 36(14), 1141-1150.
Huang, W., Chu, H., Dong, B., Hu, M. & Yu, Y. (2015). A membrane combined process
to cope with algae blooms in water. Desalination, 355, 99-109. doi:10.1016/
j.desal.2014.09.037.
An Evaluation of the Performance of the Coagulation/Flocculation … 215
Sun, F., Pei, H. Y., Hu, W. R., Li, X. Q., Ma, C. X. & Pei, R. T. (2013). The cell damage
of Microcystis aeruginosa in PACl coagulation and floc storage processes.
Separation and Purification Technology, 115, 123-128.
Sun, F., Pei, H. Y., Hu, W. R. & Ma, C. X. (2012). The lysis of Microcystis aeruginosa in
AlCl 3 coagulation and sedimentation processes. Chemical Engineering Journal,
193, 196-202.
USEPA. (2009). Drinking Water Contaminant Candidate List 3-Final Federal Register,
74(194), 51850-51862.
Wang, H. Q., Mao, T. G., Xi, B. D., Zhang, L. Y. & Zhou, Q. H. (2015). KMnO4 pre-
oxidation for Microcystis aeruginosa removal by a low dosage of flocculant.
Ecological engineering, 81, 298-300. doi:10. 1016/j.ecoleng.2015.04.015.
Westrick, J. A., Szlag, D. C., Southwell, B. J. & Sinclair, J. (2010). A review of
cyanobacteria and cyanotoxins removal/inactivation in drinking water treatment.
Analytical and bioanalytical chemistry, 397(5), 1705-1714.
WHO. (2003). Cyanobacterial Toxins: Microcystin-LR Guidelines for Drinking-Water
Quality. World Health Organization Geneva, pp. 95–110.
Wu, C. D., Xu, X. J., Liang, J. L., Wang, Q., Dong, Q. & Liang, W. L. (2011). Enhanced
coagulation for treating slightly polluted algae-containing surface water combining
polyaluminum chloride (PAC) with diatomite. Desalination, 279(1-3), 140-145.
doi:10.1016/ j.desal.2011.06.007.
Zamyadi, A., Coral, L. A., Barbeau, B., Dorner, S., Lapolli, F. R. & Prévost, M. (2015).
Fate of toxic cyanobacterial genera from natural bloom events during ozonation.
Water research, 73, 204-215.
Zhang, X., Fan, L. & Roddick, F. A. (2014). Feedwater coagulation to mitigate the
fouling of a ceramic MF membrane caused by soluble algal organic matter.
Separation and Purification Technology, 133, 221-226.
Zhang, Y., Tian, J., Nan, J., Gao, S., Liang, H., Wang, M. & Li, G. (2011). Effect of PAC
addition on immersed ultrafiltration for the treatment of algal-rich water. Journal of
Hazardous Materials, 186(2), 1415-1424.
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 12
ABSTRACT
In Morocco, like all the countries of the world, socio-economic activities coupled
with population growth and changes in consumption patterns generate a significant
production of municipal solid waste. This is accompanied by a significant increase in
leachates produced during the fermentation of waste. These leachates have a considerable
impact on the environment. Moroccan regulations require leachate treatment to reduce
environmental impacts. Several techniques are currently used for the decontamination of
leachate such as (coagulation flocculation, biological treatment, filtration, oxidation, ...).
*
Corresponding Author Email: [email protected].
218 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Indeed, the majority of the physicochemical treatments used for the purification of
the leachates intervenes as pretreatment or finishing stage to complete the treatment
chain, or to eliminate a specific pollutant.
The addition of a polymer as a flocculant improves the effectiveness of coagulation
by inorganic coagulants. In fact, these flocculants make it possible to obtain large flocs,
which settle rapidly and which resist shear forces (Bratby, 2006).
According to Bratby (2006), when a liquid is already destabilized after the addition
of an inorganic coagulant, the polymers increase the rate of orthokinetic flocculation by
the formation of larger flocs. This is done through a sufficient absorption affinity between
the polymers and the flocs.
The choice of polymers must be made judiciously because their effectiveness
depends on several parameters, namely: the nature of the polymer, the pH, the
temperature, the type of inorganic coagulant, the nature and size of the pollutant to be
eliminated, the concentration of the coagulant and flocculant (Zemaitaitiene et al., 2003,
Bolto and Gregory, 2007).
The first part of this work concerns the coagulation flocculation of young leachate
from the Kenitra city landfill. The tests were carried out by adding ferric chloride mixed
with three flocculants, namely: the chitosan, the Superfloc SD2065 and the Himoloc.
The dose of coagulant was fixed at 6 g/L (determined from preliminary tests), while
the doses of flocculants ranged from 3,3 to 20 mL/L. The evaluation of treatment
efficiency as well as the optimal doses of the flocculants was determined by the
monitoring of the removal of turbidity, COD, phenol, color, Absorbance at 254 nm and
the volume of sludge collected after 24 hours.
The purpose of the second part reveals the effect of the order of introduction of the
reagents on the efficiency of leachate treatment by coagulation flocculation.
The reagents used are ferric chloride and the three flocculants previously mentioned.
We varied the order of introduction by adding:
The effect of the order of introduction of the reagents was evaluated in terms of
elimination of turbidity, COD, color, Abs at 254 nm, phenol, detergents, ammonium ions,
nitrite, nitrate and total phosphorus.
INTRODUCTION
Solid waste management remains an issue for countries around the world. In
Morocco, the improvement of the standard of living, as well as the change in production
and consumption patterns have led to an increase in the volume of solid waste, estimated
at 6.9 million tonnes/year according to the third report on the state of of Morocco’s
environment, published in 2015. The quantity produced is of the order of 0.76 kilos per
day per inhabitant in urban areas and 0.3 kilograms per day per inhabitant in rural areas.
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 219
This increase affects the quality of the collection and generates a saturation of
landfills. They become a source of contamination for soil, air and human health, instead
of being the ultimate solution for getting rid of waste.
Landfills are sources of methane, which is a greenhouse gas, and toxic liquid
effluents called “leachate” (Renou et al., 2008). The design of a landfill must ensure the
capture of biogas for energy recovery, and the drainage and treatment of leachates before
their release into the natural environment.
In this context, Morocco is part of a proactive approach to waste management, whose
main objective is to improve waste collection, sorting and treatment conditions while
limiting the nuisance caused by methane emissions and leachate production.
The genesis of leachates is the result of a very complex process. Indeed, as soon as
they are deposited, the waste is subject to biodegradation processes that are the result of a
multitude of biological and physicochemical reactions. The water that infiltrates the
layers of waste is the main vector of this biodegradation. Thus, leachate results from the
solubilization of compounds during the percolation of rainwater through the pile of
waste, as well as moisture contained in the waste itself (Souabi et al., 2010). The
composition of leachate depends on several parameters including: The age of the landfill,
the type of waste landfilled and the technique used for landfilling (Chofqi et al., 2007).
The challenge is now significant because it will be necessary to ensure effective
drainage of leachates and their treatment for reuse or rejection in the receiving
environment. This approach will provide a significant water resource for use, particularly
in areas experiencing water shortages and drought.
Several treatment techniques are implemented to reduce and eliminate the pollution
contained in leachates. These techniques are physicochemical, biological or a
combination of both.
Among these methods, we distinguish the technique of coagulation flocculation.
Several researchers have used this process for the pre or post treatment of landfill
leachate, to improve its biodegradability or to eliminate refractory organic matter (Aziz et
al., 2007).
The goal is to eliminate many types of pollutants, namely: COD, BOD5, NH4+,
Phosphorus, color, humic and fulvic compounds, phenol ... (Maranon et al., 2008,
Poznyak et al., 2008; al., 2014).
In a study conducted by Zheng et al. (2009), coagulation flocculation using
Aluminum chloride shown to be an effective treatment for the removal of phthalic acid
esters. The removal of these pollutants reached 32% and 50% respectively for young and
intermediate leachates.
Li et al. (2010) compared the effect of four coagulants, namely: Alumina sulphate
(Al2(SO4)3), ferric chloride (FeCl3), Aluminum polychloride and polyferric Sulphate, for
the treatment of stabilized leachate. At the end of this study, Polyferric Sulfate gave the
220 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
best results by eliminating 93% of suspended solids, 97% of turbidity, 70% of COD and
74% of toxicity. The minimum volume of sludge produced was 32 mL/L.
Abood et al. (2014) studied the effectiveness of coagulation flocculation as a
pretreatment of stabilized leachate from the Chang Shankou landfill in China. They were
able to remove 70,6% of the COD, 49,4% of the BOD5 and 26,6% of NH3-N by adding
1,2 g/L of polyferric Sulphate and 1 mL of 0,1% polyacrilamide flocculant at an optimal
pH of 5.
The objective of this chapter is to determine the type and the dose of the flocculant
suitable for the leachate treatment of Ouled Berjal landfill (Situated at Kenitra City in
Morocco), and to study the effect of the order of introduction of the coagulant and the
flocculant on the effectiveness of coagulation flocculation.
The goal of coagulation flocculation is to remove suspended solids (SS) and colloids
contained in a liquid (Adamczyk, Z. 2003), which are collected as flocs and removed by
settling or flotation.
This process takes place in two stages:
Coagulants Used
There are three types of coagulants: Iron or aluminum-based metals Salts, organic
coagulants and bio-coagulants (Renou el al., 2008).
Mineral Coagulants
In general, metal salts based on iron or aluminum are the most commonly used.
Among these salts are: Ferric chloride (FeCl3), Ferric sulphate (Fe2(SO4)3) and ferrous
sulphate (FeSO4), Aluminum sulphate (Al2(SO4)3), Aluminum chloride (AlCl3) and
Sodium Aluinate (NaAlO2) (Maranon et al., 2008, Liu et al., 2012). The effectiveness of
these salts is due to the hydroxides (resulting from their hydrolysis in the medium) whose
solubility is strongly related to pH.
Figure 1 shows the coagulation diagrams of iron as a function of pH. We notice that
the iron coagulation diagram is divided into three zones:
The third zone is that in which the surface of the colloids is modified until the
zeta potential of the particles is reversed, which can lead to the
stabilization of these colloids or the destabilization of the formed flocs.
Flocculation coagulation is no longer effective in this range of pH and iron
concentration.
Organic Coagulants
The use of organic coagulants is more effective when the waters to be treated are
composed of fine suspended matter. They are synthetic cationic polyelectrolytes such as:
melamine-formaldehyde, epichlorohydrin DimethylAmine (EPI-DMA) of chemical
formula C5H12ONCl, and PolyDiAllylDimethylAmmoniumChloride (POLYDADMAC)
of chemical formula C8H16NCl.
According to Mouchet (2000), these coagulants have the advantage of not modifying
the pH of the medium, reducing the volume of sludge produced and also reducing the
dose of reagent necessary for optimum coagulation (less reagent is added).
Bio-Coagulants
Although metal salts are the most used, their negative effect on the environment and
human health is undeniable. As a result, the orientation towards the use of coagulants of
natural origin has become an extreme necessity (Chaibakhsh et al., 2014).
These types of coagulants are known by their biodegradability, low cost and
especially their non-toxicity to humans and the environment. They can be used alone or
mixed with other conventional coagulants (Nwaiwu and Bello, 2011, Awad et al., 2013).
Patel and Vashi (2012) conducted a comparative study of three types of
biocoagulants: Moringa oleifera seeds, corn seeds and chitosan for the treatment of
wastewater from the textile industry. The best results in terms of removal of COD and
color were obtained by the first bio-coagulant, whit respective removal efficiencies of
74,1% 58,5%.
Moreover, Rasool et al. (2016) showed the effectiveness of basil seeds (dried,
crushed and then sieved) mixed with Alumina Sulphate and Aluminum Polychloride for
the pretreatment of landfill leachate. The optimal conditions obtained by the
experimental design methodology are: 15min contact time, neutral pH, and 1:1 alumina
sulphate: Basil ratio. Under these conditions, 64.4% of the COD and 77.8% of the color
were eliminated.
Floculants Used
flocculants make it possible to obtain large flocs, which settle rapidly and which resist
shear forces (Bratby, 2006).
According to this author, when a medium is already destabilized following the
addition of an inorganic coagulant, the polymers increase the rate of orthokinetic
flocculation by the formation of larger flocs. This is done through a sufficient absorption
affinity between the polymers and the flocs.
The choice of polymers must be made judiciously because their effectiveness
depends on several parameters, namely: the nature of the polymer, the pH, the
temperature, the type of inorganic coagulant, the nature and the size of the pollutant to be
eliminated (Zemaitaitiene et al., 2003, Bolto and Gregory, 2007).
These flocculants may be inorganic or organic. For mineral flocculants, activated
silica is the most used.
Organic flocculants can be of natural origin such as alginates which are extracted
from marine algae (Degrémont, 2005) or synthetic such as polyelectrolytes. These are
either anionic, cationic or nonionic.
They are used to improve the elimination of organic matter (Tatsi et al., 2003) and
also promote the agglomeration of flocs and improve their settling (Renou et al., 2008).
Solberg and Wagberg (2003) have shown that the removal efficiency of a pollutant
depends on the type of polymer, its surface charge density, the concentration and
solubility of the polymer, the chemical affinity of the polymer in thr surface, ionic
strength....
Nowadays, the use of biopolymers is very important because of their ecological
nature. According to Renault et al. (2009), there is no risk of secondary contamination
because the biopolymers do not leave residual metals either in the treated water or in the
sludge produced.
In this study, leachate was sampled from Ouled Berjal landfill located in Kenitra city
in Morocco.
Kenitra is a town of 76Km² located 45Km north of Rabat. It is one of the most
important industrial centers in Morocco (the fourth largest industrial city in the country).
The climate in this city is sub humid to semi-arid with an average rainfall of about
600 mm. According to the urban commune of Kenitra, the population reached in 2014
nearly 1,061,435 habitants.
The rapid evolution of the city of Kenitra is accompanied by increased production of
solid waste, which quantity was estimated at 120,000 tons in 2011.
224 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Leachate is stored in four basins noted B1, B2, B3 and B4 (Figure 3). The samples
were collected in 30L plastic bottles, and then they were transported to the laboratory,
stored at 4°C and subsequently characterized.
The Ouled Berjal landfill covers an area of 20 ha and receives an average of 329 tons
per day of waste. The landfill has:
The leachates collected were analyzed for many parameters, namely: pH,
conductivity, turbidity, Absorbance at 254nm, color, Chemical Oxygen Demand (COD),
Biochemical Oxygen Demand (BOD5), NH4+, NO3-, phenol, total phosphorus, surfactants,
suspended matter.
Turbidity was measured using a turbidimeter Model 2100N HACH, color and
absorbance at 254 nm with a UV/Visible spectrophotometer Model 9200 and pH by a pH
meter model 6209. Conductivity was measured using a Conductivity pH meter model
YK-2001PH.
The suspended matter was determined by the centrifugation method.
The chemical oxygen demand (COD) was determined using Open Reflux Method
(5220-B). The BOD5 is analyzed according to the manometric method using the
WARBURG respirometer.
For the analysis of surfactants: The anionic surfactants form, with methyl violet, a
complex soluble in toluene and which can be determined by the colorimetric method. The
absorbance of the residue recovered after extraction is spectrophotometrically read at a
wavelength of 615 nm.
The ammonium ions were assayed according to the method stated by Rodier in 1996,
which consists in the treatment of these ions with a solution of sodium nitroprusside. The
spectrophotometric reading is then carried out at a wavelength of 630 nm.
Determination of nitrates was carried out in accordance with the NF EN ISO 78-90
January 1997 (T 90-045) method.
226 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
The total phosphorus was determined according to the NF T 90-023 January 1997
method.
Phenol was measured using the Folin-Ciocalteu reagent method (Macheix et al.,
1990). This reagent develops, in the presence of phenolic compounds, a blue color whose
intensity is read by a spectrophotometer at a wavelength of 725 nm.
Reagents Used
Ferric Chloride
Table 2 shows the physicochemical characteristics of the FeCl3 coagulant.
Polyacrylamide Himoloc
Himoloc DR3000 flocculant is a cationic polyacrylamide of medium molecular
weight, specially adapted for the physico-chemical treatment of wastewater because it
contains a special molecule that facilitates the separation of the colloidal material by
flotation.
This flocculant is soluble in any type of wastewater, regardless of its quality.
The table below lists the main characteristics of the Himoloc DR3000 flocculant.
Chitosan
Chitosan is a substance not widespread in nature. It exists only in the cell walls of
certain fungal microorganisms such as zygomycetes fungi and in mycelium.
Figure 4 shows the chemical structure of chitosan.
The particular properties of this biopolymer are related to the presence of the amine
function carried by the carbon 2. These properties depend strongly on the pH:
At acidic pHs, chitosan carries several positive charges and can bind to negatively
charged molecules. It is considered a good coagulant or flocculant (Fang et al., 2001);
At pH’s exceeding neutral, chitosan loses its positive charges while releasing the
electronic doublets of nitrogen. Thus, chitosan is considered an excellent complexing
agent, particularly heavy metals (Yen et al., 2009).
Parameter Value
FeCl3 (% by mass) 39,0 – 41,0
Fe3+ (% by mass) 13,4 – 14,2
Density at 20°C (kg/dm3) 1,400 – 1,440
pH at 20°C <1
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 227
Parameter Value
Density (g/cm3) 1,2
Viscosity (cP) < 600
pH 3,0 – 4,1
Parameter Value
Density (g/cm3) 1,01 - 1,05
Viscosity (cP) 270
The leachate flow rate was measured using a submersible pump with an hourly flow
rate of 20 m3/h and operating automatically with floats. We were able to determine the
daily flow rate of the leachate collected by a daily recording of the operating time of the
pump.
The monthly flow of the leachate admits a maximum value of 2900 m3/month and a
minimum value of 640 m3/month. Indeed, maximum flows are presented during the
months of January, February and March, which coincides with the rainy season of the
year. During this period, stormwater increases the volume of the leachate produced,
causing overflows in the storage basins.
Figure 5. Variation of the monthly and daily flow of leachate produced in Ouled
Berjal landfill.
While the months of July and August present the minimum flows, as it is the summer
period known by relatively high temperatures, leading to the evaporation of a large part
of the stored leachates.
The characterization of leachate aims to evaluate the risk that the landfill may
represent to the environment.
We carried out 3 sampling campaigns during the months of February, March (Rainy
season) and October (End of summer) 2015.
Figures 6, 7 and 8 report the results of the characterization.
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 229
Variation of pH
The leachates stored in the different basins B1, B2 et B3 have a basic pH varying
between 8 and 9. The pH values are strongly related to the existence of the volatile
organic compounds in the leachate. Indeed, during the first phase of waste fermentation
(Acidogenesis phase), leachates contain high levels of volatile compounds, thus inducing
low values of pH, which are generally less than 4. Over time, the landfill ages and the
leachates are depleted of volatile organic compounds, resulting in increased pH values up
to 7 or higher (Kjeldsen et al., 2002).
During the decomposition, waste is subject to chemical and biological reactions
leading to the production of ammonium ions and carbonic acid. The dissociation of the
latter results in the formation of hydrogen cations and bicarbonate anions which have a
direct effect on the pH of the percolation water or leachate. Thus, the pH of the leachate
is strongly influenced by the nature of the dissolved materials existing in the medium. An
acidic pH typically characterizes a young leachate containing a large organic charge
(Naveen et al., 2017). Moreover, an alkaline pH is an indicator of leachate maturity,
which is found to be depleted in organic matter and rich in refractory to biodegradation,
such as humic substances (Renou et al., 2008).
BOD5 (mg/L)
Conductivity
COD (mg/L)
NH4+ (mg/L)
phosphorus
NO3-(mg/L)
Surfactant
Turbidity
(FD=100)
(mS/cm)
SS (g/L)
(FD=20)
(mg/L)
(mg/L)
(mg/L)
Phenol
254nm
Abs at
(NTU)
Color
Total
pH
Basin
B1 8,4 22,1 114,5 0,272 0,300 0,7 4519,2 992,5 1126,6 24,5 24,1 626,4 374
B2 8,7 20,5 73,6 0,292 0,340 0,8 4412,8 895,0 1122,4 46,0 14,0 393,8 382
B3 8,7 22,2 72,7 0,324 0,308 1,0 4372,2 899,3 1109,2 68,1 16,7 230,3 334
Figure 6. Variation of COD and BOD5 contained in leachates from basins B1, B2 et B3 during the three campaigns.
Figure 7. Variation of pH, conductivity, turbidity and SS contained in leachates from basins B1, B2 et B3 during the three campaigns.
Figure 8. Variation of color, Abs at 254nm, phenol and surfactant contained in leachates from basins B1, B2 et B3 during the
three campaigns.
Figure 9. Variation of phosphorus, NH4+ and NO3- contained in leachates from basins B1, B2 et B3 during the three campaigns.
234 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
In our case, the BOD5/COD ratio ranges from 0,14 to 0,3, indicating that the
leachates in the three basins are intermediate leachates with mean biodegradability.
However, it is possible to envisage a biological treatment for these effluents (Renou et
al., 2008).
Landfills
Parameter
(Morocco)
(Morocco)
(Morocco)
(Morocco)
(Algeria)5
ElKerma
El jadida
(China)6
Beijing
Agadir
Rabat
Fez
1
4
pH 8,4 8,84 7,11 4,3 8,19 8
COD (mg/L) 14 900 920 53 999,6 72 000 19 333 8528
BOD5 (mg/L) 8700 55,33 20 000 44 000 3301 5669
NH4+ (mg/L) 3740 114 2,41 0,10 2726 1154
NO3- (mg/L) - 2,37 11,3 62 0,92 2,21
Total Phosphorus (mg/L) 49,73 - - 188,4 0,60 5,2
Turbidity (NTU) 1619 - - - 732 -
Conductivity (mS/cm) 22,8 27,18 - 14,7 120,3 -
1
Bakraouy et al. (2016)
2
Chofqi et al. (2004)
3
Ezzoubi et al. (2010)
4
Jirou et al. (2014)
5
Bennama et al. (2010)
6
Zhu et al. (2013)
Variation of Conductivity
The conductivity values obtained ranged from 38 to 13,5 mS/cm, with an average of
22 mS/cm. It approaches the values recorded by Bakraouy et al. (2016) which is 22,8
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 235
mS/cm, Muller et al. (2015) which is 30 mS/cm, as well as that of Chofqi et al. (2007)
which is 26,5 mS/cm. we notice that these values are quite high, showing that the
leachates of the three basins are highly mineralized. Moreover, the conductivity reaches
low values during the winter season, which can be explained by the dilution of the
leachate by stormwater.
While during the summer season, the conductivity values are high, which can be
related to the bacterial activity that is favored by the increase in temperature. This
increase stimulates enzymatic reactions, thereby releasing high amounts of mineral
elements such as chlorides. Khattabi et al. (2002) demonstrated that conductivity is
closely related to the content of chloride ions.
Variation of Phenol
Phenol usually comes from resins, paints, plastics, sand molding foundries. In
addition, phenol is the most important benzene derivative, after styrene. Ozonation,
chlorination, coagulation-flocculation, or adsorption on activated carbon (Aravindhan et
al., 2009) have been developed for its elimination due to its toxicity. It is considered to be
an inhibitor of biomass in biological treatments.
We note that the concentration of phenol admits a maximum value of 1258 mg/L and
a minimum value of 945,8 mg/L. These values are higher than those reported by Silva et
al. (2013), which reaches 162 mg/L, as well as that recorded by Bannama et al. (2010),
which reaches 37 mg/L.
It should be noted that the concentration of phenol increases over time, since
anaerobic storage promotes the degradation of phenolic derivatives while producing
phenol.
Variation of Surfactants
For the three basins, the analysis of surfactants revealed moderate values ranging
from 18 to 77,6 mg/L.
According to Wang et al. (2004), surfactants come from various domestic, urban and
industrial activities (production of detergents and fertilizers, pharmaceutical and
petrochemical industries, etc.). They are major inhibitors of the development of biomass
promoter of biodegradability. Borghi et al. (2011) explained that when the chain length of
the alkyl group increases, the hydrophobicity of the molecule increases also, inducing
high toxicity. On the other hand, a high number of ethylene oxide units reduces the
hydrophobicity of the molecule and induces a decrease in toxicity.
other organisms, causing the imbalance of the aquatic ecosystem. This phenomenon is
called: Eutrophication. The elimination of phosphorus is therefore one of the major
environmental issues.
The values obtained during the sampling campaigns ranged from 9,3 to 34,4 mg/L.
These values exceed those reported by Silva et al. (2013), with an average of 19,1 mg/L,
and by Zhu et al. (2013), with an average of 5,2 mg/ L, but still much lower than that of
the Agadir landfill, which reached 188,4 mg/L (Jirou et al., 2014).
3
NH4+ + O2 → NO−
2 + H2 O + 2H
+
2
1
NO− −
2 + O2 → NO3
2
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 237
Moreover, the elimination of these two elements (nitrates and nitrites) is commonly
called denitrification, it is carried out under anoxic conditions. This reduction in nitrogen
gas is governed by the reaction below:
NO− −
3 → NO2 → NO → N2 O → N2
Variation of Nitrates
Nitrates result from the oxidation of nitrogen, they represent the most common
nitrogenous form in natural waters. According to Rodier et al. (2009), nitrates represent a
major part of mineral nitrogen, which together with organic nitrogen constitute the total
nitrogen. Toxicity of nitrates has been widely studied by Karine et al. (2003).
Indeed, nitrates alter the quality of water as well as human health. On the one hand,
the reduction of nitrates to nitrites favors the formation of methemoglobin which causes
the oxidation of iron II to iron III, which makes it incapable of transporting oxygen. On
the other hand, nitrates contribute to the endogenous synthesis of N-nitroso compounds
which are strongly linked to the risk of cancerous diseases.
Nitrate levels in the three basins ranged from 172 to 592 mg/L. These values remain
below the maximum value determined by Souabi et al. (2011) for leachates from the city
of mohammedia in Morocco, which reaches 845 mg/L.
In a research conducted by Tahiri et al. (2014), the characterization of the leachate
from the city of Meknes reveals a concentration of nitrates about 751,1 mg/L, which
exceeds that revealed by the present study. On the other hand, the nitrate concentration is
lower in the leachate of the city of El jadida, 2,4 mg/L (Chofqi et al., 2007), of Fez
citywhich reaches 11,3 mg/L (Ezzoubi et al., 2010) and of Oran city which reaches 0,92
mg/L (Bennama et al., 2010).
Throughout Figure 3, we notice that the nitrate concentration is higher in the summer
season, due to the nitrification of ammonium ions and their conversion to nitrates. The
latter are accumulated in the leachate, since the denitrification is carried out under
optimum temperature conditions ranging from 35 to 50°C.
238 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Table 7 reports the physicochemical characteristics of the leachate used in this study.
The leachate studied has a blackish color, a fecal odor and an average turbidity of
180 NTU. Organic load expressed in terms of COD reaches 21 g/L, which shows that this
leachate is young and biodegradable. This concentration is higher than that reported by
Chofqi et al. (2007), which was 990 mg/L for the leachate from the El Jadida landfill.
Furthermore, we note high concentrations of nitrogen pollutants and moderate phenol
content (96,2 mg/L) and surfactants (47,1 mg/L). These concentrations are much lower
than those detected by Bakraouy et al. (2016) for the Rabat landfill, being 241,8 mg/L for
phenol and 58 mg/L for surfactants.
The metal charge contained in the leachate from the Ouled Berjal landfill is typical of
a domestic landfill since the concentration of metallic elements is similar to that detected
in other household landfills. According to Christensen et al. (2001), heavy metals
composition of a household landfill is as follows:
Chromium comes from cardboard paper and wood that is collected with household
waste. Its content is 190 μg/L, it is much lower than the concentration detected in the
leachate of the Mohammedia landfill (Chtioui et al. 2005).
Parameter Value
Turbidity (NTU) 180
COD (mg/L) 21120
Phenol (mg/L) 96,2
NH4+ (mg/L) 536,9
NO3- (mg/L) 456,0
NO2- (mg/L) 760,0
Total phosphorus (mg/L) 31,5
Surfactant (mg/L) 47,1
Abs 254 nm (FD = 200) 0,258
Color (FD = 20) 0,569
Hg (µg/L) 6,5
Cr (µg/L) 190
Cd (µg/L) 25
Zn (µg/L) 85
Pb (µg/L) 85
Cu (µg/L) 45
*
FD: Dilution factor
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 239
This part concerns the study of the coagulation flocculation of young leachate. These
tests were carried out by the addition of ferric chloride mixed with three flocculants,
namely: chitosan, Superfloc SD2065 and Himoloc.
The dose of the coagulant was set at 6g/L (determined from preliminary tests), while
flocculant doses ranged from 3,3 to 20 mL/L. The evaluation of the quality of the
treatment was determined by monitoring removal efficiencies of: Turbidity, COD,
phenol, color, Abs at 254 nm and the volume of sludge collected after 24 hours.
Figure 9 illustrates the variation of these parameters as a function of different
concentrations of the flocculants.
Referring to Figure 10 (a), Abs at 254 nm decreases progressively with the addition
of the different flocculants until reaching a minimum value, and then increases following
an excess of flocculant.
The removal efficiencies of Abs at 254 nm obtained by the addition of the three
flocculants vary from 75 to 85%. Superfloc SD2065 flocculant and chitosan achieved a
removal of around 83%. In the case of Himoloc flocculant, the highest removal was 81%
at a concentration of 20 mL/L.
The results we obtained remain superior to those obtained by Ntampou et al. (2006),
with a removal of 75% of Abs at 254nm, adding 365 mg/L of coagulant.
Tzoupanos et al. (2008) conducted coagulation flocculation tests via a new coagulant
named PSI. Removal efficiencies of Abs at 254 nm obtained varied from 15 to 70%. The
best removal was 75%, obtained by the addition of 300 mg/L of PACSi 1/15 coagulant.
These authors show that Alumina sulphate is the least suitable coagulant for the
elimination of Abs at 254 nm.
Elimination of Phenol
Figure 10 (b) shows the effectiveness of ferric chloride mixed with the three
flocculants in terms of phenol removal. The maximum removals were achieved using
Superfloc SD2065 and chitosan, and are respectively 99 and 88%. The optimal dose of
both flocculants is 13 mL/L. For Himoloc flocculant, phenol was reduced by 81% by
adding 3 mL/L.
242 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Achour and Guesbaya (2005) conducted CF testing for the removal of phenolic
organic compounds and humic substances. Low phenol removal was achieved ranging
from 9,26 to 32%. The authors concluded that phenolic structures are little affected by
Coagulation flocculation. Krou (2010) showed that charcoal adsorption is an effective
method for phenol removal. The disadvantage of this technique is the high cost generated
by the regeneration of coal. Indeed, the pollutant remains concentrated on the surface of
the solid which itself becomes a waste that must be transported and treated within
specialized sites.
Elimination of COD
We note that COD removal increases with the increase in the concentration of the
three flocculant. However, beyond the optimal concentration, COD elimination
decreases. This is due to the excessive adsorption of polymer on the colloidal surfaces,
which induces the production of stabilized colloids. These colloids are positively charged
and cause electrostatic repulsion of suspended particles (Abu Hassan et al., 2009).
The most effective flocculant is Superfloc SD2065, with a maximum removal of
around 86% at a concentration of 13 mL/L. In addition, a dose of 10 mL/L of Himoloc
flocculant leaded to an elimination of 82% of COD initially contained in the leachate.
COD was reduced by 73% by adding 17mL/L of chitosan. Several researchers use ferric
chloride as coagulant for the treatment of landfill leachate. This is the case of Tatsi et al.
(2003), who studied the treatment of stabilized leachate from the Thessaloniki landfill in
Greece. Without pH adjustment, the addition of 1,5 g/L of FeCl3 + eliminated 80% of
COD, while a dose of 1,5 g/L of Al3+ leaded to an elimination around 38%. These authors
also tested the effect of several flocculants: Anionic (A321), cationic (K1370 and K506)
and neutral (N200) on the treatment of young and partially stabilized leachates. They note
that in the case of partially stabilized leachates, the addition of these flocculants has no
effect on improving the reduction of organic matter. In addition, for young leachates, the
elimination of organic pollution was improved by the simultaneous effect of pH
adjustment (to a value of 10) and the addition of flocculants. As a result, 80% of COD
was removed by the addition of flocculant A321 and ferric chloride with a concentration
of 1,5 g/L (Tatsi et al., 2003).
Volume of Sludge
According to Assou et al. (2015), the volume of sludge produced strongly depends on
physicochemical characteristics of the leachate as well as the effectiveness of
Coagulation flocculation treatment.
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 243
Based on Figure 10 (d), we note that the mixture of ferric chloride with Superfloc
SD2065 flocculant induces the formation of less sludge (after 24 hours of settling),
compared to the other two flocculants. Indeed, the minimum volume of sludge is 43% for
the mixture FeCl3+Superfloc SD2065, 46% for FeCl3+Himoloc and 50% for
FeCl3+chitosan.
At low concentrations, FeCl3+chitosan mixture produces less sludge. According to
Renault et al. (2009), chitosan significantly increases the density of sludge and facilitates
drying. In addition, chitosan is a biodegradable biopolymer, which allows sludge
degradation by microorganisms.
Removal efficiencies obtained slightly exceed those achieved by Maranon et al.
(2008), which reached 32 and 30% respectively at 30 min and 24h decantation.
Assou et al. (2015) noted that the volume of sludge produced increases dramatically
as the concentration of flocculants increases. According to these authors, this may be due
to the cationic nature of the flocculants used in the study. These flocculants have a high
molecular weight, which promotes bond formation between small flocs that have the
ability to capture colloidal particles and promote their settling.
These same authors add that organic polymers generally produce less sludge than
inorganic salts. Consequently, sludges resulting from coagulation via ferric chloride
mixed with a polyelectrolyte are compact.
Elimination of Turbidity
Figure 10 (e) shows that turbidity removal is little affected by the change in
concentration of different flocculants.
Indeed, removal efficiencies are of the order of:
Elimination of Color
Figure 10 (f) shows that color removal is little influenced by the flocculant
concentration. The maximum values reached:
Conclusion
Comparing the results obtained following the addition of the various flocculants, we
note that the mixture of ferric chloride with Himoloc is better suited to the treatment of
leachate from ouled Berjal landfill with a minimum volume of 10 mL/L (Figure 11).
The purpose of this section is to reveal the effect of order of introduction of the
reagents on the efficiency of treatment of landfill leachate by coagulation flocculation.
In this context, leachate was tested for coagulation flocculation by ferric chloride and
three different flocculants, namely: Chitosan, Superfloc SD2065 and Himoloc.
The optimal doses were determined following preliminary tests, and are as follows:
The first reagent is added to the sample and stirred at 150 rpm for 10 minutes;
The second reagent is added and stirred at the same speed for 10 min;
The sample is then subjected to slow stirring at 30 rpm for 30 min;
The supernatant is removed after 1 hour of decantation.
The effect of the order of introduction of the reagents was evaluated in terms of
elimination of turbidity, COD, color, Abs at 254 nm, phenol, surfactant, ammonium ions,
nitrites, nitrates and total phosphorus.
246 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Elimination of Turbidity
The results reported in Figure 12 (a) show that the introduction of coagulant followed
by flocculant provides the best turbidity removal efficiencies.
Turbidity was reduced by 90% by chitosan and Himoloc, and 87% by Superfloc
SD2065.
In a study conducted by El Bada et al. (2010), doses of lime, calcite and soda,
ranging from 0 to 10 g/L were added to leachate from the Azemmour landfill to eliminate
turbidity, COD and metals. In terms of turbidity, lime was the most effective in
eliminating 80%.
Abu Hassan et al. (2009) studied coagulation flocculation of wastewater from textile
industry using chitosan. The results showed that the reduction in turbidity and COD were
respectively 94,9 and 72,5% following the addition of 30 mg/L of chitosan at
a pH of 4.
Elimination of COD
COD removal is maximum when the coagulant and the flocculant are added
simultaneously. COD was reduced by 86% by mixing the ferric chloride with the
Himoloc flocculant, and 82% by the other two flocculants.
These efficiencies exceed those obtained by other researchers (Zeng et al., 2008, El
bada et al., 2010, Ez-zoubi et al., 2010).
Figure 12. Effect of the order of introduction of the reagents on removal of pollutants in the leachate.
Precipitation by lime, NaOH and calcite has been studied by El bada et al. (2010).
They showed that the COD decreases with the increase in the concentration of the three
coagulants. Its elimination reached 31, 24 and 16 respectively for lime, NaOH and
calcite.
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 249
Ez-zoubi et al. (2010) obtained a maximum reduction of 35% in terms of COD. The
authors showed that COD removed corresponds to the hard COD and its removal is due
to the adsorption on ferric chlorides (Wang et al., 2002).
Elimination of Color
From Figure 12 (c), there is little difference between the different flocculants as well
as between the different orders of introduction. Indeed, color removal exceeded 97%.
Rasool et al. (2016) used experimental design methodology for the
optimization treatment of landfill leachate using basil seed powder as a biocoagulant.
Optimal conditions for the three factors are a contact time of 15 min, a pH of 7 and a
Alum:Basil ratio of 1:1. COD and color were eliminated by 64,4% and 77,8%
respectively
Syafalni et al. (2012) evaluated the treatment efficiency of landfill leachate
with lateritic soil as coagulant. An optimal dose of 14 g/L at a pH of 2 makes it
possible to reduce the COD by 65,7%, color by 81,8% and the ammoniacal nitrogen by
41,2%.
Zahrim et al. (2010) performed coagulation flocculation of highly concentrated dye
solutions (Acid Black 210). The authors compared the effect of five flocculants, mixed
with aluminum sulphate used as a coagulant. The best removal efficiency of color (37%)
was obtained by adding polyDADMAC flocculant.
Elimination of Phosphorus
Moreover, in the study of Jirou et al. (2014), the Agadir landfill leachate was
subjected to continuous aeration for 60 days. Total phosphorus was reduced by 92,1% at
the end of the test, while in the control sample the removal was only 51,2%.
Khattabi et al. (2002) studied elimination of the total phosphorus contained in
leachate from Etueffont landfill by natural lagooning. The authors achieved a reduction
of 70% in summer and 14% in winter. The low rate reached in the winter season was
explained by the decrease in algal assimilation of phosphorus as well as the re-solution of
this element, which is favored by a low pH (of the order of 4).
Volume of Sludge
Phenol and detergents are major inhibitors of the development of biomass promoting
biodegradability. Their elimination is therefore of great necessity.
From Figure 12, the introduction of chitosan and Superfloc SD2065 flocculant first
improves detergent removal, with removals of 64 and 57%, respectively. The lowest
removal (23%) was obtained by Himoloc.
Unlike phenol which elimination is favored by the introduction of the coagulant first,
specifically for chitosan (64%) and Himoloc.
Figure 13 reports the effect of the order of introduction of the reagents on the
removal of nitrogen compounds.
According to Rodier et al. (1996), mineral nitrogen composed of ammonia, nitrates
and nitrites, constitutes the major part of the total nitrogen. Elimination of nitrogen
compounds is of great importance in the field of water treatment, given their negative
effect on the environment as well as on human health.
According to Umapriya and Shrihari (2010), elimination of ammoniacal nitrogen is
probably done by the adsorption of ammonium ions on the coagulant. Thanks to their
positive charge, the ammonium ions act as complementary species to improve the effect
of Al3+ and Fe3+ cations on the coagulation of the suspended matter, thus acting bridging
elements as follows:
We note that the maximum removals in terms of ammonium ions are quite low. They
are of the order of 14, 46 and 39% respectively for chitosan, Superfloc SD2065 and
Himoloc.
Orders of introduction leading to these results differ from one flocculant to another.
Indeed, for chitosan, the most suitable is to introduce the flocculant first. While for the
other two flocculants, it is advisable to introduce the coagulant first.
For nitrites, elimination efficiencies reached:
64% for chitosan and 83% for Himoloc by introducing the flocculant first;
74% for Superfloc SD2065 by introducing reagents at the same time.
252 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
In addition, the elimination of nitrates is favored by adding the coagulant first, for
Superfloc SD2065 (99%) and Himoloc (62%). While adding ferric chloride first followed
by chitosan reduced the nitrate content by 55%.
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 253
CONCLUSION
In this chapter, we studied the treatment of leachate from ouled Berjal discharge by
flocculation coagulation.
For young leachate from Ouled Berjal landfill, we evaluated the effectiveness
of ferric chloride as coagulant mixed with various flocculants, such as:
Chitosan, Himoloc DR3000 and Superfloc SD2065. The best results were
obtained by mixing the coagulant with Himoloc DR3000 polymer, with removal
efficiencies of 90, 86, 99, 83 and 99,4% respectively for turbidity, COD, phenol, Abs at
254nm and color. Optimum doses are 6g/L for the coagulant and 13,3 mL/L for the
flocculant.
We also examined the effect of order of introduction of reagents on the
effectiveness of coagulation floculation. Indeed, we concluded that the elimination of
turbidity, ammonium ions, Abs at 254 nm, phenol, total phosphorus and
nitrates is favored by the introduction of the coagulant first. Minimal volume
of sludge and high COD removal are achieved by simultaneous introduction
254 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
ACKNOWLEDGMENTS
This study was conducted as part of the research and development program initiated
by the State Secretariat for Sustainable Development.
REFERENCES
Abood, AR; Bao, J; Du, J; Zheng, D; Luo, Y. Non biodegradable landfill leachate
treatment by combined process of agitation, coagulation, SBR and filtration, Waste
Management, 34, pp. 439–447, (2014).
Achour, S; Guesbaya, N. Coagulation-floculation par le Sulfate d’Aluminium de
composés organiques phénoliques et de substances humiques, larhyss journal, N°4,
pp. 153-168, (2005). [Coagulation-flocculation with aluminum sulphate of phenolic
organic compounds and humic substances]
Adamczyk, Z. Particle adsorption and deposition: role of electrostatic interactions.
Advances in Colloid and Interface Science, 100-102, pp. 267-347, (2003).
Ahmad, AL; Sumathi, S; Hameed, BH. Coagulation of Residue Oil and Suspended Solid
in Palm Oil Mill Effluent by Chitosan, Alum and PAC, Chemical Engineering
Journal, 118, pp. 99-105, (2006).
Ariffin, A; Shatat, RSA; Nik Norulaini, AR; Mohd Omar, AK. Synthetic Polyelectrolytes
of Varying Charge Densities but Similar Molar Mass Based on Acrylamide and Their
Applications on Palm Oil Mill Effluent Treatment, Desalination, 173, pp. 201-208,
(2005).
Assou, M; Madinzi, A; Anouzla, A; Aboulhassan, MA; Souabi, S; Hafidi, M. Reducing
Pollution of Stabilized Landfill Leachate by Mixing of Coagulants and Flocculants: a
Comparative Study, Int j eng and inn tech., 4, pp. 20-24, (2014).
Awad, M; Wang, H; Li, F. Preliminary study on combined use of Moringa seeds extract
and PAC for water treatment, Res. J. Recent Sci., 2, pp. 52–55, (2013).
Aziz, HA; Alias, S; Adlan, MN; Asaari, FAH; Zahari, MS. Colour removal from landfill
leachate by coagulation and flocculation processes, Bioresource Technol., 98, pp.
218–220, (2007).
Bakraouy, H; Abouri, M; Souabi, S; Digua, K; Yaacoubi, A; Jada, A. Optimization of
Fresh Landfill Leachate Treatment by a Response Surface Methodology, J. Colloid
Sci. Biotechnol., 5, pp. 182–189, (2016).
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 255
Bashir, MJK; Abdul Aziz, H; Yusoff, MS; Adlan, MN. Application of response surface
methodology (RSM) for optimization of ammoniacal nitrogen removal from semi-
aerobic landfill leachate using ion exchange resin, Desalination, 254, pp. 154–161,
(2010).
Bennama, T; Younsi, A; Derriche, Z; Debab, A. Caractérisation et traitement physico-
chimique des lixiviats de la décharge publique d’El-Kerma (Algérie) par adsorption
en discontinu sur de la sciure de bois naturelle et activée chimiquement, Water Qual.
Res. J. Can., 45 (1), pp. 81–90, (2010). [Characterization and physicochemical
treatment of leachates of the El-Kerma dump (Algeria) by batch adsorption on
natural and chemically activated sawdust]
Bolto, B; Gregory, J. Organic polyelectrolytes in water treatment. Water Res., 41 (11),
pp. 2301–2324, (2007).
Bratby, J. Coagulation and Flocculation in Water and Wastewater Treatment, second ed.,
IWA Publishing, London., (2006).
Chaibakhsh, N; Ahmadi, N; Zanjanchi, MA. Use of Plantago major L. as anatural
coagulant for optimized decolorization of dye-containing wastewater, Ind. Crops
Prod., 61, pp. 169–175, (2014).
Chi, FH; Cheng, WP. Use of chitosan as coagulant to treat wastewater from milk
processing plant. J Polym Environ, 14, pp. 411–7, (2006).
Chofqi, A; Younsi, A; Lhadi, EK; Mania, J; Mudry, J; Veron, A. Lixiviat de la décharge
publique d’El Jadida (Maroc): Caractérisation et étude d’impact sur la nappe
phréatique, Déchets sciences et techniques, N°46, pp. 4-10, (2007). [Leachate of the
El Jadida landfill (Morocco): Characterization and impact study on groundwater,
Waste science and technology]
Divakaran, R; Pillai, VNS. Flocculation of kaolinite suspensions in water by chitosan,
Water Res, 35, pp. 3904–8, (2001).
El bada, N; Assobhei, O; Kebbabi, A; Mhamdi, R; Mountadar, M. Caractérisation et
prétraitement du lixiviat de la décharge de la ville d’Azemmour. Déchets, Sciences et
Techniques - Revue francophone d’écologie industrielle, N°58, pp. 30-36, (2010).
[Characterization and pre-treatment of the leachate of the landfill of the city of
Azemmour. Waste, Science and Technology - Francophone Review of Industrial
Ecology]
Ez zoubi, Y; Merzouki, M; Bennani, L; El Ouali Lalami, A; Benlemlih, M. Procédé pour
la réduction de la charge polluante du lixiviat de la décharge contrôlée de la ville de
Fès, Déchets, sciences et techniques - Revue francophone d’écologie industrielle, 58,
pp. 22-29, (2010). [Process for the reduction of the pollution load of the leachate of
the controlled landfill of the city of Fes, Waste, science and technology - French-
speaking magazine of industrial ecology]
Jirou, Y; Harrouni, MC; Belattar, M; Fatmi, M; Daoud, S. Traitement des lixiviats de la
décharge contrôlée du Grand Agadir par aération intensive, Rev. Mar. Sci. Agron.
256 Hajar Bakraouy, Salah Souabi, Khalid Digua et al.
Vét., 2 (2), pp. 59-69, (2014). [Treatment of leachate from the Grand Agadir
controlled landfill by intensive aeration]
Khalil, F; Bouaouine, O; Chtioui, H; Souabi, S; Aboulhassan, MA; Ouammou, A.
Traitement des lixiviats de décharge par coagulation-floculation, [Treatment of
leachate from discharge by coagulation-flocculation] J. Mater. Environ. Sci., 6 (5),
pp. 1337-1342, (2015).
Krou, NJ. Thèse de doctorat. Université de Toulouse. Etude expérimentale et
modélisation d’un procédé séquentiel AD-OX d’élimination de polluants organiques,
(2010). [Experimental study and modeling of a sequential AD-OX process for the
removal of organic pollutants]
Li, W; Hua, T; Zhou, QX; Zhang, SG; Li, FX. Treatment of stabilized landfill leachate by
the combined process of coagulation/flocculation and powder activated carbon
adsorption, Desalination, 264, pp. 56–62. (2010).
Liu, X; Li, XM; Yang, Q; Yue, X; Shen, TT; Zheng, W; Luo, K; Sun, YH; Zeng, GM.
Landfill leachate pretreatment by coagulation-flocculation process using iron-based
coagulants: Optimization by response methodology, Chem Eng J., 200-202, pp. 39–
51, (2012).
Mahmud, K; Hossain, MD; Shams, S. Different treatment strategies for highly polluted
landfill leachate in developing countries, Waste Management, 32 (11), pp. 2096-
2105, (2012).
Maranon, E; Castrillon, L; Fernandez-Nava, Y; Fernandez-Mendez, A; Fernandez-
Sanchez, A. Coagulation–flocculation as a pretreatment process at a landfill leachate
nitrification–denitrification plant, J. Hazard. Mater., 156, pp. 538–544, (2008).
Martin, MA; Gonzalez, I; Berrios, M; Siles, JA; Martin, A. Optimization of coagulation–
flocculation process for wastewater derived from sauce manufacturing using factorial
design of experiments, Chemical Engineering Journal, 172, pp. 771– 782, (2011).
Millot, N. Les lixiviats de décharge contrôlée. Caractérisation analytique. Etude des
filières de traitement. [Landfill leachate controlled. Analytical characterization.
Study of the treatment sectors.] Thèse de doctorat INSA Lyon, (1986).
Mouchet, P. Traitement des eaux avant utilisation. Matières particulaires. Techniques
d’ingénieur, G, 1, 170, (2000). [Water treatment before use. Particulate matter.
Engineering techniques]
Muller, GT; Giacobbo, A; Chiaramonte, EADS; Rodrigues, MAS; Meneguzzi, A;
Bernardes, AM. The effect of sanitary landfill leachate aging on the biological
treatment and assessment of photoelectrooxidation as a pretreatment process, waste
Management, 36, pp. 177-183, (2015).
Naveen, BP; Mahapatra, DM; Sitharam, TG; Sivapullaiah, PV; Ramachandra, TV.
Physico-chemical and biological characterization of urban municipal landfill
leachate., Volume 220, Part A, pp. 1–12, (2017).
Treatment of Leachate from the Ouled Berjal Landfill in Morocco … 257
Chapter 13
ABSTRACT
Natural and assisted particles sedimentation are crucial operations during the
treatment of wastewater generated in many industrial processes, above all in the
wastewater treatment field. This operation is usually used as an effluent pre-treatment;
normally essential for reducing suspended solids, organic load, turbidity and colour so as
to achieve the correct clarification of wastewater. This chapter introduce the coagulation
and flocculation concepts as well as the influence of operational conditions on assisted
sedimentation.
At industrial level, the factual use of both processes has been proved on different
wastewater. In this sense, this chapter outline researches about combining or comparing
assisted sedimentation (coagulation, electrocoagulation, neutralization, etc.) with other
operations such as oxidation processes in order to evaluate the solids removal of the
complete designed wastewater treatment focusing on OMW treatment. In addition, in this
work new trends about multifunctional and novel eco-friendly bio-degradable flocculants
and coagulants were reviewed and suggested for OMW pretreatment.
*
Corresponding Author Email: [email protected].
260 Gassan Hodaifa and Cristina Agabo García
1. INTRODUCTION
In the last century, two-phases production process (TP-PP) has become to the most
efficient and eco-friendly olive oil production extraction method. As the most used in
Spanish mills, TP-PP applies a centrifugation step after common operations: olive
washing, crushing and malaxing. This operation involve the use of horizontal centrifuge
‘decanter’ with two exists: one for crude olive oil which contain impurities and the other
for a wet solid residue named in Spain ‘alperujo.’
In this context, olive oil mill wastewater (OMW) is a result of the mixture of
different wastewaters generated during olive oil production process. In other words, it is
the sum of wastewater from washing olives (WOW), vegetation water, wastewater from
olive oil washing (WOOW) and others wastewaters such as additional water added to
malaxing system, cleaning devices wastewaters, etc. (Dermeche et al., 2013; Garcia-Ivars
et al., 2015).
On average, olives washing machine needs 1 m3 of drinking water per each
tonne of olives. This wastewater contains basically natural organic matter (NOM) coming
from sources of natural waters and the impurities such as dust, little stones,
microorganisms or even herbicides and pesticides (Guardia-Rubio et al., 2006 and 2007).
Avoiding water addition in malaxing system, TP-PP generates a 0,33-0,35 m3 of
wastewater per tonne of olives and bringing tannins, pectins, lignines, long fatty acids,
reduced carbohydrates, proteins and bacterial and plants growth inhibitors
phenolic compounds (Dermeche et al., 2013; Bouknana et al., 2014; Gerasopoulos et al.,
2015).
Regarding OMW treatment, natural or assisted particles sedimentation is a
fundamental step to be used as effluent pretreatment. This basic operation contribute to
the treatment of wastewater by reducing its suspended solids, organic load, turbidity and
colour. Since the presence of different kinds of suspended solids in olive oil mills
wastewater (OMW), colloids are charged stabilized particles that avoid aggregation
leading to non-sedimentable particles.
Coagulation and flocculation techniques consist of using chemical agents that reduce
the charge of the particles and promote micro-particles formation. It is commonly named
coagulation and flocculation indistinctly. However, on the one hand, coagulation
comprises the electrostatic destabilization of the colloids that form microflocs. But on the
other hand, flocculation is the flocs growth by chemical bonds formation resulting in
sedimentable cluster of particles (Grant et al., 2001; Martínez-Nieto et al., 2011; Rytwo
et al., 2013).
OMW Pretreatment by Assisted Sedimentation Methods 261
1. Charge neutralization.
2. Entrapment mainly by Van der Waals forces.
3. Adsorption forces.
4. Complexation with coagulant metal ions into insoluble particulate aggregates.
where ‘Rcol’ is the rate of particles collision and ‘Rbr’ is the rate of flocs breakage. ‘α’ is
the collision efficiency factor that means the number of effective collisions which result
in attachment among the total of collisions. This factor is not constant, while particles
sizes are increasing, the number of particles in the system is reduced leading to a decrease
in efficiency of collision (α·Rcol). When both parts of the equation have the same value,
the net rate of flocs growth is zero. At this point, the steady-state is reached and the size
of the flocs have got its maximum level. It is important to take account that floc breakage
may not be readily re-formed in some irreversible breakages and the ‘α’ parameter would
be equal to 0 (Gregory and Dupont, 2001).
Floc stability analysis in suspension is possible by determining not only the strength
but also the number of the bonds holding the floc together. The floc strength can be
262 Gassan Hodaifa and Cristina Agabo García
a) Collision rate, that depends on the number and the size of the colloidal particles,
as well as, mixing conditions (timing and velocity), of course surface, and size of
the decanter (Grant et al., 2004; Matilainen et al., 2010). In this sense, the
behavior of floc size increasing thorough the time in each aggregate is related to
the rest of particles aggregations by an infinite number of non-lineal differential
equations.
b) Particles interaction. Attractive and repulsive forces between particles are due to
surface charges. This also involves the physico-chemical characteristic of
effluent particles such as size, functionality, charge and hydrophobicity.
Normally, organic matter with high molecular mass needs less coagulant dosage due
to that the mechanism of removal is, basically, neutralization of charges. In addition,
OMW Pretreatment by Assisted Sedimentation Methods 263
2.2. pH Effect
products of ferric chloride (normally used as coagulant) are cationic so interaction with
negative colloids is easy under correct pH value. Fe+3 is arounded by six molecules of
water which will be progessively replaced by hydroxyl ions due to polarization and loss
of protons because of pH values (Sahu and Chaudhari, 2013).
Figure 1. Chemical equilibrium diagram of iron species presence in function of pH value at initial iron
concentration = 10 M. The diagram was obained by using Medusa-Hydra chemical equilibrium
software.
Salt concentrartions have also a crucial influence due to the supply of divalent cations
and destabilizing anions (bicarbonate, chloride, and sulfate). As an example, Ferric
chloride, ferric sulphate and alum are acidic, leading to lower pH condition after
application (Yan et al., 2008). Moreover, alkalinity and pH are related. In water with
lower alkalinity, coagulant addition may consume all the available alkalinity, depressing
the pH to values favourable for coagulation. In contrast, at higher alkalinity conditions it
is necessary the addition of great amount of coagulant in order to decrease pH values.
Floc size, strength and re-construction of flocs after shear break-up during
flocculation process are influenced by temperature due to changes in coagulant
chemistry. Temperature affects the solubility of the metal hydroxide precipitate and the
rate of formation of the metal hydrolysis products. In addition, temperature conditions
also affect the viscosity of clarified water after flocculation.
Gregory et al., (2004) compared aluminium sulphate, ferric sulphate and
polyaluminium chloride products (PACL). It was concluded that warmer temperatures
OMW Pretreatment by Assisted Sedimentation Methods 265
(>15ºC) generally produce bigger flocs that not only are easily to be broken but also
tough to be reformed. On the other hand, lower temperatures favoured lower size flocs
which suffer less number of breakages and recover their structure better. Moreover, the
variation of size and strength of flocs depends on the type of flocculant used. As an
example, the results of this study showed that aluminium based coagulants produce flocs
more sensible to temperature variation than the ones formed using ferric sulphate. In
addition, flocculation size also depends on the dosage of flocculant. At lower aluminium
dosage (1.4 mg/L), the size of flocs had more variation than at higher dosage (3.4 mg/L)
in the case of using PACL. In many cases, as lower temperatures favour
flocculation/coagulation process, it is better wait for the cooling of warm water before
starting the coagulation step. In full-scale plants, cold temperatures tend to reduce
kinetics of hydrolysis reactions and particle flocculation formation due to a decrease of
alum and PACL solubilities and an increase in water viscosity. For this reason, higher
coagulant dosages and additional flocculation time are required at a low temperature
(Frank et al., 2000).
less amounts of problems with final disposal of sludge. This fact is considered as one of
the advantage of using potassium ferrate among inorganic salts.
On the other hand, among polymers, it can be found inorganic polymeric coagulants
and synthetic and natural organic polymeric materials that are used as flocculants (dos
Santos et al., 2018). Polymers can be cationic, anionic or non-ionic (neutrally charged).
By and large, polymers show low diffusion rates raising the viscosity of solution. For this
reason, it is compulsory, a great mechanical dispersion of the polymer into the water, as it
mentioned before, by applying a strong first step of mixing in a level that polymers can
not be degraded (Sahu and Chaudhari, 2013).
Since the beginning of the last decade, inorganic polymerized flocculants (IPFs) have
been increasingly used in water treatment not only in Europe and Japan but also in North
America owing to availability and reduced price. IPFs are formed by the addition of
anions to conventional coagulant salts acquiring new coagulant properties. IPFs are
considered to be improved in particulation, organic removal, lower alkalinity
consumption and lesser sludge production (Sinha et al., 2004; Tzoupanos and Zouboulis,
2011). As an example, Van Benschoten and Edzwald (1990) compared the alum with and
prepared PACL. Results claimed that PACL showed a shape based on small spheres (<25
mm) and/or chain-like structures, whereas alum appeared as a fluffy and porous flocs
(ranging from 25 to 100 mm). The advantages of the structures formed by PACL
included: lesser turbidity in clarified water and less sensibility to temperature changes.
These benefits could arise from differences in their hydrolysis, nucleation, growth, and
aggregation leading to a new structure formation which interact more efficiently with
more fractions of organic matter (Van Benschoten and Edzwald, 1990). Others such as
polymeric ferric sulphate (PFS) are being used not only because it is superior capability
in removing turbidity, algae, colour and natural organic matter (NOM) but also avoiding
corrosion (Shi et al., 2004).
It is well known the advantage of synthetic polyelectrolytes compared to natural
ones. In a study conducted by Aguilar et al., (2003) some polymers were used as a
coagulant aids with ferric sulphate in liquid effluent from a slaughterhouse. Results
showed that, using anionic polyacrylamide or polyvinyl alcohol polymers increase the
total particulates removal from 80 to 99 and 93%, respectively (Aguilar et al., 2003).
However, organic polyelectrolytes generally present toxicity due to its monomer presence
in clarified water. In fact, some countries have already forbidden or restricted the use of
some polyelectrolytes for drinking water applications (WHO, 2000). For this reason,
there is a growing interest in replacing synthetic polyelectrolytes with natural and
sustainable alternatives, such as combination of PAM with activated starch (Lapointe and
Barbeau, 2017).
Natural organic coagulants presents a minimal health risk to living organisms, and
are highly biodegradable compared to inorganic coagulants. In this sense, exist a large list
of natural coagulant–flocculants that come from vegetable, animal, or microorganism
OMW Pretreatment by Assisted Sedimentation Methods 267
sources used in the treatment of natural water and wastewater, such as Nirmali seed,
maize mesquite, species of Opuntia, chitosan and xanthan (Choumane et al., 2017). As an
example, herbal coagulants (mustard seed extract, guar gum, Moringa oleifera) have
proven effective for treatment of drinking water and wastewater (Sanghi et al., 2006). In
addition, they could be more efficient that inorganic coagulants. In other study conducted
by dos-Santos et al., (2018) three tannin-based coagulants (Tanfloc, Acquapol C1 and
S5T) were compared to aluminium sulphate in order to treat starch effluent. Conclusions
showed that the highest colour and turbidity removal occurred for natural coagulants
Acquapol S5T and Tanfloc SL. at 320 mg/L dos-Santos et al., (2018). Sinha et al., (2004)
proposed a strategy to select the most effective coagulant by three steps:
modify pH after treatments such as FeCl3 that reduce pH values to 2-3 favouring AOPs
after flocculation (Papaphilippou et al., 2013). Ginos et al., (2006) also have used a
combination of two agents: i) 5000 mg/L of Fe(II) or 30,000 mg/L of lime (inorganic
coagulants) and ii) 287 mg/L FO-4700-SH (cationic flocculant). The conclusions calimed
that 230 mg/L cationic polyelectrolyte is more effective alone than in combination,
reducing TSS 97.4%, total phenolic compounds (TPCS) 49.7% and COD 17.1%.
However, the amount of sludge produced during the combined use of inorganic coagulant
and poly-electrolyte was lower than that during separation with the poly-electrolyte
alone. Fe(II) and polyelectrolyte did not change pH conditions but the use of lime made
the sample alkaline. But combined process using lime always resulted in greater TPCs
removal than treatment with Fe(II) and therefore it is reduced the toxicity of samples.
Considering the restriction in using organic polycations, before mentioned, it is
important to decide which process is better according to the searched criteria: more
quality, flocculant more eco-friendly, less sludge amount formed, final pH conditions,
etc. In addition, velocity, as well as, volume sludge formation during flocculantion is also
important. In this sense, Martínez-Nieto et al., (2011) studied kinetic equation of
flocculation operation after OMW treatment by Fenton. In this research, it was registered
volumes and velocities of the formed sludge using different concentrations of several
coagulants/flocculants. In the same study Nalco-77171 reached the maximum volumes of
formed sludge and clarified water, 13,5% v/v and 86,5% v/v, respectively. Furthermore,
the main quality parameters of wastewater were determined obtaining %DQOremoval =
18.3% and %Ironremoval = 99.6%. Hodaifa et al., (2015) proposed a treatement using
flocculation as a pretreatment before a foto-Fenton operation. Results showed that the
flocculant with the major capability of sludge formation (Nalco-9913, 3% v/v) was not
the flocculant with the major capability of reducing organic load (QG-2001, DQOremoval =
20.4%).
In spite of high efficiencies in using charged organic polymeric flocculants in OMW,
others authors have tested other kinds of coagulant/flocculant with similar results
chitosan, starch, alum and ferric chloride (Meyssami and Kasaeian, 2005). Even
neutralization as an assited sedimentation operation was tested as pretreatments. Mert et
al., (2010) proposed two pretreatments before Fenton operation: i) Chemical pre-
treatment based on acid cracking and coagulation–flocculation which achieve 67%
CODremoval and ii) Fenton process with >80% COD removal.
Different authors evaluate organic removal capacity of electrocoagulation method. In
this case, the reaction of charged ionics species from wastewater and
coagulants/flocculants improve remotion of organic load. In this sense, Hande and
Arslan-Alaton (2014) used acid cracking (at pH 2.0 and T = 70 °C) and filtration as
pretreatments. Pretreated wastewater was subsequently treated by different techniques: i)
Coagulation with FeCl3, ii) Ca(OH)2 precipitation, iii) electrocoagulation using stainless
steel electrodes and iv) Fenton process as chemical treatment. Among them,
270 Gassan Hodaifa and Cristina Agabo García
4. NEW TRENDS
Water desinfection.
Degradation of organic and inorganic contaminants in advanced oxidation
process (AOPs).
Elimination of suspended particles and heavy metals in water.
Ferric chloride salt also has multiple functions: acting as a catalyst in AOPs and as an
inorganic coagulant during sedimentation. However, this chemical mineral present
corrosive effect in pipes and pumps. On the other hand, Ferrate (VI) is a strong oxidizer,
which in acidic condition, has the highest oxidation potential even beyond ozone
contributing to AOPs and therefore, disinfection. In addition, it is not corrosive and has a
high solubility in water (approximately 15 g/L) where is converted into trivalent iron or
ferric hydroxide (Fe(OH)3) which is a strong coagulant. Hence, ferrate (VI) is suggested
as a novel chemical able to eliminate heavy metals, suspended particles and organic
matter, without producing hazardous by-products. In any case, more researchs at large-
scale must be conducted in order to study this treatment deeply (Chen et al., 2015;
Talaiekhozani et al., 2017). Furthermore, over the last years, there is a tendency of
synthetizing composite coagulants or using natural biodegradable flocculants. Preparation
of composite coagulants consist of adding organic or inorganic modifiers to the
respective coagulants in order to improve coagulation-flocculation efficiency by adding a
high cationic charge. As an example Al–Fe composite unify advantages of both Al and
Fe polymers, to overcome the problems associated with the individual coagulants. In this
sense, other metals have been incorporated to aluminium ferric polymers such as Zn and
Si (Zhu et al., 2012; Wei et al., 2016). Jia et al., (2017) used composite coagulant
polymeric aluminum ferric sulfate (PAFS) in order to treat urban wastewater where
optimal conditions were [PAFS] = 45 mg/L and pH = 8.5. The removal %COD and
%turbidity were 82.8% and 98.2%, respectively. However, the presence of sulphates or
sodium ions (which could be sometimes found in OMW) could interfere into coagulation
(Jia et al., 2017). On the other hand, some researchers have used natural and eco-friendly
flocculating agents such as ones based on cellulose or derived from plants (Khiari et al.,
OMW Pretreatment by Assisted Sedimentation Methods 271
2010; Zhu et al., 2010; Nourani et al., 2016; Kono, 2017; Ajao et al., 2018) to reduce the
toxicity and the treatment costs. However, no studies were carried out in OMW.
Considering these new trends, new studies of applicability of these novel flocculants
must be conducted in order to improve flocculation pretreatment of OMW.
REFERENCES
Achak, M., Ouazzani, N., Yaacoubi, A. and Mandi, L. (2008). Modern olive mill effluent
characterization and their treatment by coagulation–flocculation using lime and
aluminium sulphate. Revue des sciences de l’eau “Journal of Water Science,” 2153-
2167.
Aguilar, M. I., Saez, J., Llorens, M., Soler, A. and Ortuno, J. F. (2003). Microscopic
observation of particle reduction in slaughterhouse wastewater by coagulation–
flocculation using ferric sulphate as coagulant and different coagulant aids. Water
Res., 37(9): 2233-2241.
Ajao, V., Bruning, H., Rijnaarts, H. and Temmink, H. (2018). Natural flocculants from
fresh and saline wastewater: comparative properties and flocculation performances.
Chem. Eng. J., 349: 622-632.
Bouknana, D., Hammouti, B., Salghi, R., Jodeh, S., Zarrouk, A., Warad, I., Aouniti, A.
and Sbaa, M. (2014). Physicochemical Characterization of Olive Oil Mill
Wastewaters in the eastern region of Morocco. J. Mater. Environ. Sci., 5: 1039-1058.
Chen, Y., Xiong, Y., Wang, Z., Chen, Y., Chen, G. and Liu, Z. (2015). UV/ferrate (VI)
oxidation of profenofos: efficiency and mechanism. Desalination Water Treat.,
55(2): 506-513.
Chen, W., Feng, Q., Zhang, G., Li, L. and Jin, S. (2017). Effect of energy input on
flocculation process and flotation performance of fine scheelite using sodium oleate.
Miner. Eng., 112: 27-35.
Choumane, F. Z., Benguella, B., Maachou, B. and Saadi, N. (2017). Valorisation of a
bioflocculant and hydroxyapatites as coagulation-flocculation adjuvants in
wastewater treatment of the steppe in the wilaya of Saida (Algeria). Ecol. Eng.,
107:152-159.
Dermeche, S., Nadour, M., Larroche, C., Moulti-Mati, F. and Michaud, P. (2013). Olive
mill wastes: Biochemical characterizations and valorization. Process Biochem., 48:
1532–1552.
dos-Santos, J. D., Veit, M. T., Juchen, P. T., da Cunha Gonçalves, G., Palácio, S. M. and
Fagundes-Klen, M. (2018). Use of different coagulants for cassava processing
wastewater treatment. J. Environ. Chem. Eng., 6(2):1821-1827.
272 Gassan Hodaifa and Cristina Agabo García
Ersoy, B., Tosun, I., Günay, A. and Dikmen, S. (2009). Turbidity removal from
wastewater of natural stone processing by coagulation/flocculation methods. Clean –
Soil, Air, Water, 37(3): 225–232.
Fitzpatrick, C. S. B., Fradin, E. and Gregory, J. (2004). Temperature effects on
flocculation, using different coagulants. Water Sci. Technol., 50(12): 171-175.
Frank, M., Gersonde, R., Rutgers van der Loeff, M. M., Bohrmann, G., Nurnberg, C.,
Kubik, P. W., Suter, M. and Mangini, A. (2000). Similar glacial and interglacial
export bioproductivity in the Atlantic sector of the Southern Ocean: Multiproxy
evidence and implications for atmospheric CO2. Paleoceanography, 15: 642–658.
Garcia-Ivars, J., Iborra-Clar, M. I., Alcaina-Miranda, M. I., Mendoza-Roca, J. A. and
Pastor-Alcaniz, L. (2015). Treatment of table olive processing wastewaters using
novel photomodified ultrafiltration membranes as first step for recovering phenolic
compounds. J. Hazard. Mater., 290: 51-59.
Gerasopoulos, K., Stagos, D., Kokkas, S., Petrotos, K., Kantas, D., Goulas, P. and
Kouretas, D. (2015). Feed supplemented with byproducts from olive oil mill
wastewater processing increases antioxidant capacity in broiler chickens. Food
Chem. Toxicol., 82: 42-49.
Gernjak, W., Maldonado, M. I., Malato, S., Caceres, J., Krutzler, T., Glaser, A. and
Bauer, R. (2004). Pilot-plant treatment of olive mill wastewater (OMW) by solar
TiO2 photocatalysis and solar photo-Fenton. Sol. Energy, 77(5): 567-572.
Ginos, A., Manios, T. and Mantzavinos, D. (2006). Treatment of olive mill effluents by
coagulation–flocculation–hydrogen peroxide oxidation and effect on phytotoxicity. J.
Hazard. Mater., 133(1-3):135-142.
Grant, S. B., Kim, J. H. and Poor, C. (2001). Kinetic theories for the coagulation and
sedimentation of particles. J. Coll. Interface Sci., 238(2): 238–250.
Gregory, J. and Dupont, V. (2001). Properties of flocs produced by water treatment
coagulants. Water Sci. Technol., 44(10): 231–236.
Gregory, J., (2004). Monitoring floc formation and breakage. Water Sci. Technol.,
50(12): 163-170.
Guardia-Rubio, M., Ruiz-Medina, A., Molina-Díaz, A. and Fernández de Córdova, M. L.
(2006). Determination of pesticides in washing waters of olive processing by gas
chromatography-tandem mass spectrometry. J. Sep. Sci., 29(11): 1578-86.
Guardia-Rubio, M., Ayora-Canada, M. J. and Ruiz-Medina, A. (2007). Effect of Washing
on Pesticide Residues in Olives. J. Food Sci., 72(2): 139-143.
Hande, B. and Arslan-Alaton, I. (2014). Treatment of olive mill wastewater by chemical
processes: effect of acid cracking pretreatment. Water Sci. Technol., 69(7): 1453-
1461.
Hodaifa, G., Páez, J. A., Agabo, C., Ramos, E., Gutiérrez, J. C. and Rosal, A. (2015).
Flocculation on the treatment of olive oil mill wastewater: Pre-treatment. Inter. J.
Chem. Mol. Eng., 9(5): 633-63.
OMW Pretreatment by Assisted Sedimentation Methods 273
Hodaifa, G., Páez, J. A., Agabo, C., Ramos, E., Gutiérrez, J. C. and Rosal, A. (2015).
Flocculation on the treatment of olive oil mill wastewater: Pretreatment. World
Academy of Science, Eng. Technol., Inter. J. Chem., Mol. Nuclear, Mater. Metallur.
Eng., 9(5): 645-650. https://waset.org/publications/10001929/flocculation-on-the-
treatment-of-olive-oil-mill-wastewater-pretreatment.
Hogg, R. (2013). Bridging flocculation by polymers. KONA Powder and Particle J., 30:
3-14.
Jaouani, A., Vanthournhout, M. and Penninckx, M. J. (2005). Olive oil mill wastewater
purification by combination of coagulation-flocculation and biological treatments.
Environ. Technol., 26(6): 633-642.
Jarvis, P., Jefferson, B., Gregory J. and Parsons, S. A. (2005). A review of floc strength
and breakage. Water Res., 39 (14): 3121-3137.
Jia, D., Li, M., Liu, G., Wu, P., Yang, J., Li, Y., Zhong, S. and Xu, W. (2017). Effect of
basicity and sodium ions on stability of polymeric ferric sulfate as coagulants.
Colloid Surface A., 512: 111-117.
Khiari, R., Dridi-Dhaouadi, S., Aguir, C. and Mhenni, M. F. (2010). Experimental
evaluation of eco-friendly flocculants prepared from date palm rachis. J. Environ.
Sci., 22(10): 1539-1543.
Kono, H. (2017). Cationic flocculants derived from native cellulose: Preparation,
biodegradability, and removal of dyes in aqueous solution. Resource-Efficient
Technol., 3(1): 55-63.
Korshin, G., Chow, C. W., Fabris, R. and Drikas, M. (2009). Absorbance spectroscopy-
based examination of effects of coagulation on the reactivity of fractions of natural
organic matter with varying apparent molecular weights. Water Res., 43(6): 1541-
1548.
Lapointe, M. and Barbeau, B. (2017). Dual starch–polyacrylamide polymer system for
improved flocculation. Water Res., 124: 202-209.
Martínez-Nieto, L., Hodaifa, G., Rodríguez, S., Giménez, J. A. and Ochando, J. (2011).
Flocculation–Sedimentation Combined with Chemical Oxidation Process. Clean –
Soil, Air, Water, 39 (10): 949–955.
Matilainen, A., Vepsäläinen, M., Sillanpää M. (2010). Natural organic matter removal by
coagulation during drinking water treatment: a review. Adv. Coll. Interface, 159: 189-
197.
Mert, B. K., Yonar, T., Kilic, M. Y. and Kestioglu, K. (2010). Pre-treatment studies on
olive oil mill effluent using physicochemical, Fenton and Fenton-like oxidations
processes. J. Hazard. Mater., 174: 122–128.
Meyssami, B. and Kasaeian, A. B. (2005). Use of coagulants in treatment of olive oil
wastewater model solutions by induced air flotation. Bioresource Technol., 96(3):
303-307.
274 Gassan Hodaifa and Cristina Agabo García
Michael, I., Panagi, A., Ioannou, L. A., Frontistis, Z. and Fatta-Kassinos, D. (2014).
Utilizing solar energy for the purification of olive mill wastewater using a pilot-scale
photocatalytic reactor after coagulation-flocculation. Water Res., 60:28-40.
Murcia, J. J., Hernández-Laverde, M., Rojas, H., Muñoz, E., Navío, J. A. and Hidalgo,
M. C. (2018). Study of the effectiveness of the flocculation-photocatalysis in the
treatment of wastewater coming from dairy industries. J. Photoch. Photobiol. A.,
358: 256-264.
Nassar, N. N., Arar, L. A., Marei, N. N., Ghanim, M. M. A., Dwekat, M. S. and Sawalha,
S. H. (2014). Treatment of olive mill based wastewater by means of magnetic
nanoparticles: Decolourization, dephenolization and COD removal. Environ.
Nanotechnol. Monitor. Manag., 1: 14-23.
Nourani, M., Baghdadi, M., Javan, M. and Bidhendi, G. N. (2016). Production of a
biodegradable flocculant from cotton and evaluation of its performance in
coagulation-flocculation of kaolin clay suspension: optimization through response
surface methodology (RSM). J. Environ. Chem. Eng., 4(2): 1996-2003.
Pala, A. and Tokat, E. (2002). Color removal from cotton textile industry wastewater in
an activated sludge system with various additives. Water Res., 36: 2920–2925.
Papaphilippou, P. C., Yiannapas, C., Politi, M., Daskalaki, V. M., Michael, C.,
Kalogerakis, N., Mantzavinos, D. and Fatta-Kassinos, D. (2013). Sequential
coagulation–flocculation, solvent extraction and photo-Fenton oxidation for the
valorization and treatment of olive mill effluent. Chem Eng. J., 224: 82-88.
Rytwo, G., Lavi, R., Rytwo, Y., Monchase, H., Dultz, S. and König, T. N. (2013).
Clarification of olive mill and winery wastewater by means of clay–polymer
nanocomposites. Sci. Tot. Environ., 442: 134–142.
Sahu, O. P. and Chaudhari, P. K. (2013). Review on chemical treatment of industrial
waste water. J. Appl. Sci. Environ. Manage, 17(2): 241-257.
Sanghi, R., Bhattacharya, B. and Singh, V. (2006). Use of Cassia javahikai seed gum and
gum-g-polyacrylamide as coagulant aid for the decolorization of textile dye solutions.
Bioresource Technol., 97(10): 1259-1264.
Sarika, R., Kalogerakis, N. and Mantzavinos, D. (2005). Treatment of olive mill
effluents: Part II. Complete removal of solids by direct flocculation with poly-
electrolytes. Environ. Int., 31(2): 297-304.
Sharp, E. L., Jarvis, P., Parsons, S. A. and Jefferson, B. (2006). Impact of fractional
character on the coagulation of NOM. Coll. Surf. A., 286(1-3): 104-111.
Shi, Y., Fan, M., Brown, R. C., Sung, S. and Van Leeuwen, J. H. (2004). Comparison of
corrosivity of polymeric sulfate ferric and ferric chloride as coagulants in water
treatment. Chem. Eng. Process, 43(8): 955-964.
Sinha, S., Yoon, Y., Amy, G. and Yoon, J. (2004). Determining the effectiveness of
conventional and alternative coagulants through effective characterization schemes.
Chemosphere, 57(9): 1115-1122.
OMW Pretreatment by Assisted Sedimentation Methods 275
Chapter 14
ABSTRACT
This chapter focuses on the preparation and characterization of the chitosan based
flocculant for removal of heavy metal ion prepared from chitosan by N-acylation with
ethylenediaminetetraacetic acid monoanhydride (EDTAM). In this study, a series of
chitosan flocculants with various degree of substitution were prepared by changing the
molar ratio of EDTAM to chitosan in the preparation. The structural characterization of
the chitosan flocculants by FTIR and 13C NMR spectroscopy were performed. The newly
introduced functional group provided properties such as being a strong chelating reagent
and an amphoteric polyelectrolyte. It was found that chitosan flocculants had good water
solubility in both acidic and basic regions and precipitated in a narrow pH region, which
is close to its isoelectric point. The chitosan flocculants has an ability to remove heavy
metal ion from aqueous solution by controlling pH or initial concentration of the
flocculant, and removed Cu(II) almost completely. The flocculation mechanism was
investigated using Cu(II) and found that flocculation occurred by charge neutralization
was sensitive both pH conditions and amount of chelated Cu(II) on chitosan flocculants.
The complexation stoichiometry of Cu(II) and the EDTA residues of chitosan flocculant
*
Corresponding Author Email:[email protected].
278 Sayaka Fujita and Nobuo Sakairi
determined by UV-vis titration. As a result, the chitosan flocculants can easily remove
Cu(II) from aqueous solution by only flocculation/precipitation process, which indicated
that the chitosan flocculants could be applicable for remediation and treatment system of
wastewater.
INTRODUCTION
Among the large number of environmental pollutants, heavy metals are major water
and soil pollutants. Metals are introduced into aquatic systems by the weathering of soils
and rocks, volcanic eruptions, and a variety of human activities, such as mining,
processing, or using metal-containing materials or products. Most heavy metals are
highly toxic or carcinogenic, and tend to readily accumulate in living organisms through
the food chain. Various methods are available for the removal of heavy metal ions,
including chemical precipitation [1], ion exchange [2], adsorption [3], oxidation [4],
electrochemical treatment [5], and membrane separation [6].
As a water treatment method, flocculation is an efficient and cost-effective
technology for the removal of suspended insoluble organic and inorganic substances from
industrial and domestic wastewater. There are two major types of flocculants: inorganic
additives such as polyaluminum chloride and polyferric sulfate, and organic
macromolecules such as polyacrylamide [7]. These three commonly used flocculants
cause aggregation of dispersed particles and hydrophobic colloids into larger flocs that
can be easily settled and separated [8, 9]. They are used mainly for removal of insoluble
organic and inorganic compounds from contaminated water. As it is considered difficult
to accumulate metal ions into large flocs, flocculation has been regarded as a less
effective method for removal of metal ions.
Chitosan is a linear polysaccharide of -(1,4)-linked 2-amino-2-deoxy-D-
glucopyranose, a semi-synthetic polysaccharide made from such abundant and renewable
biomass as shells of crabs, shrimps, and insects. Behaving as a cationic polyelectrolyte in
acidic aqueous solution, chitosan and its grafted polymers have been investigated as
flocculants to reduce turbidity in wastewater. Chitosan has also the amino groups which
reactive with various transition metal ions. However, the ability of chelating with metal
ions is weaken by protonation of the amino groups, and chitosan has limitation of using
as a flocculant for removal of heavy metal ion.
The purpose of this study is to explore the properties of the chitosan based flocculant
prepared form chitosan with ethylenediaminetetraacetic acid (EDTA) (Figure 1) such as
structural characterization, removal of Cu(II), complexation stoichiometry in order to
clarify flocculation mechanism.
Chitosan Based Flocculants for the Removal of Heavy Metal Ions 279
EXPERIMENTAL
Materials
Figure 1. Scheme for chitosan based flocculant (EDTA grafted chitosan) synthesis through N-acylation
of chitosan with EDTA monoanhydride.
Apparatus
A series of six chitosan based flocculants were prepared from chitosan with EDTA.
EDTA monoanhydride was prepared by a procedure reported by Capretta et al. [11].
Chitosan (2.50 g, 15.6 mmol of D-glucosamine units) was dissolved in 2% aqueous acetic
acid-methanol (1000 ml, 1:1(v/v)). EDTA monoanhydride (0.53 g, 1.95 mmol, 0.25
molar equivalents to the glucosamine unit) was added to the solution, and then the
mixture was stirred vigorously at room temperature for 24 h. The pH of the reaction
mixture was adjusted to 8 by 2% aqueous NaHCO3, and the same amount of EDTAM
was added to the mixture. The mixture was stirred at room temperature for 12 h, and
precipitated by acidification of the mixture to pH 2-3 with 2 mol/l HCl. The resulting
precipitate was separated by filtration, washed with methanol, and then dissolved in 2%
aqueous NaHCO3. The solution was subjected to ultrafiltration using a membrane
(Advantec UK-10; Toyo Roshi Co., Ltd., Japan) with molecular weight cutoff 20,000
against deionized water, and lyophilized to give chitosan flocculant; ED-ch (Na form) 1
(2.59 g, D.S. 7.5%) as white hygroscopic amorphous powder. Using a similar method to
prepare the ED-ch, the other ED-ch 2–6 were prepared. The molar ratio of the chitosan
and EDTA for the preparation of these ED-ch are listed in Table 1.
Solubility Tests
The solubility was evaluated by turbidity measurement based on the method re-
ported by Kubota et al. [12]. Fine powder of ED-ch or chitosan (20 mg) was prepared by
dissolving in 0.1 mol/l HCl (20 ml), and adjusted the pH by addition of small amount of 3
mol/l HCl or 3 mol/l NaOH solutions. The optical transmittance of each sample was
recorded on UV-visible spectrometer at 600 nm. The zeta potential value of each sample
was also measured at different pH.
ED-ch (1.00 g) was dissolved in deionized water (50 ml) to prepare 20g/l of ED-ch
stock solution. The ED-ch stock solution was diluted with deionized water to provide
various concentrations of ED-ch solution. A Cu(II) solution (1000 ppm) prepared from
Cu(NO3)2 and 0.1 mol/l HNO3 was diluted with deionized water to give various
concentration of Cu(II) solution. One ml of the ED-ch solution was added to 19 ml of the
Cu (II) stock solution in a 50 ml centrifuge tube, and pH of the mixture was adjusted by
the addition of small amounts of 3 mol/l HNO3 or 3 mol/l CH3COONa. The mixture was
shaken at 120 rpm at room temperature for defined times. The resulting precipitate was
Chitosan Based Flocculants for the Removal of Heavy Metal Ions 281
settled by centrifugation at 4,000 rpm for 15 min and the supernatant was collected
immediately. The residual Cu(II) concentration in the supernatant was determined by
atomic absorption spectrometer.
The removal of Cu(II) with ED-ch was calculated according to the following
equation:
where C0 is the initial Cu(II) concentration and Csup is the residual Cu(II) concentration of
the supernatant. The experiments were performed three times.
For comparison, the same experiments were also performed using 20 g/l of chitosan
solution in aqueous acetic acid, which was prepared by dissolving in 1wt% acetic acid
aqueous.
Sample Molar ratio of the starting material D.S. (%)a Water solubilityb
(EDTAM:chitosan)
1 0.25:1 7.5 Insoluble
2 0.5:1 21 Insoluble
3 1:1 39 Soluble
4 1:1 45 Soluble
5 1.5:1 58 Soluble
6 3:1 79 Soluble
a
calculated from integral of H-1protons.
b
sample (5 mg) was dissolved in H2O (5 ml). Transmittance at 600 nm of the sample 95% higher than deionized
water referred as “soluble”.
to 2.0. After stirring each freshly prepared analytical sample for 30 min at room
temperature, the UV-vis spectrum was recorded between 200 and 900 nm.
Figure 2 shows the FTIR spectra of chitosan and ED-ch 1-6. These spectra were
normalized by the absorption of C-O-C in the glucopyranose at 1071 cm-1 [16]. With the
exception of the absorption at 1071 cm-1, these spectra display common adsorption bands.
The broad absorption at 3435 cm-1 corresponds to NH and O-H stretching vibration, and
absorption band between 2959 and 2888 cm-1 are attributable to the C-H stretching
vibration. The characteristic absorption bands were observed at 1653, 1592, and 1324 cm-
1
due to the C=O stretching vibration (amide I), N-H bending vibration (amide II), and C-
N stretching vibration (amide III) from the secondary amide, respectively [17]. The
absorption bands of -(1,4) glycoside bridge at 1151 and 914 cm-1. The intensity of
the bands at 1653 and 1592 cm-1 of ED-ch were obviously increased as compared to those
of chitosan, and the new absorption bands at 1636 and 1405 cm-1 due to C-O
asymmetrical and symmetrical stretching vibration of the carboxylate group was detected
in the spectrum of ED-ch [18, 19]. Furthermore, absorption band of ester (C=O stretch)
was not observed at 1720-1770 cm-1 [20, 21]. These observations suggested highly
selective formation of the amino group of chitosan by the reaction with EDTAM. In the
spectra of ED-ch 1 and 2, the appreciable alternation was not observed comparing with
that of chitosan. The spectra of 3-6 showed that intensity of the amide I and II bands at
1653 and 1592 cm-1 increased with an increase in the molar ratio of EDTAM to chitosan.
Figure 2. FTIR spectra of (a) chitosan, ED-ch (b) 1, (c) 2, (d) 3, (e) 5 and (f) 6. These spectra were
normalized by the peak intensity of the carbonyl bands at 1071 cm-1.
Figure 3 shows the 13C NMR spectra of chitosan in DCl/D2O and ED-ch 3 in
NaOD/D2O. In the spectrum of chitosan, the resonances at 98.3, 77.8, 75.6, 70.9, 61.1
284 Sayaka Fujita and Nobuo Sakairi
and 56.8 ppm were assigned to the carbonyl carbon, C1, C4, C5, C3, C6 and C2 [22]. In
the spectrum of ED-ch, the new resonances appearing at 61.6 and 55.0 ppm due to two
kinds of methylene carbon in EDTA moiety [23]. Moreover, the new resonances at 182.5
and 182.1 ppm were assignable to carbonyl carbons of carboxylate groups in the EDTA
moiety, and the signal at 177.5 ppm corresponds to carbonyl carbon of amide group.
These results also supported the assigned structure of ED-ch.
The degree of substitution (D.S.) of the EDTA residue in the products was calculated
the area ratio between the anomeric protons of substituted (H-1’) and unsubstituted (H-1)
glucosamine residues in 1H NMR spectrum (spectra not shown). The results of
calculation are shown in Table 1. The D.S. of ED-ch was able to change by the molar
ratio of starting chitosan and EDTAM used.
Figure 3. 13 C NMR spectra of chitosan in DCl/D2O (top) and ED-ch 3 (bottom) in NaOD/D2O at rt.
Figure 4. The optimal transmittance (T%) and the zeta potential of aqueous solutions of ED-ch 3 and
chitosan under various pH. Inset: The photographs of ED-ch with 0.1% aqueous solution at pH 4.0 (a)
immediately after adjusting pH and (b) after leaving to stand for 10 min.
286 Sayaka Fujita and Nobuo Sakairi
a
opened bar means “soluble”; closed bar means “insoluble”; hatched bar means “the transmittance was
unstable”.
b
determined from the zeta potential.
consideration of homogeneity of the reaction. When both solutions of ED-ch and Cu(II)
are mixed, the chelation of Cu(II) with EDTA occur quickly in the homogeneous
conditions.
Figure 5. The effect of pH on the removal of Cu(II) contacting with chitosan and ED-ch 3 by
flocculation/precipitation. Initial concentration of Cu(II) 10 mg/l, dose of polymer ligand 1 g/l, contact
time 24 h, shaking rate 120 rpm, rt.
Figure 6. The effect of contact time on the removal of Cu(II) by ED-ch 3 at pH 4.5. Insrt: The
magnified figure at range of contact time 0-30 min. Initial concentration of Cu(II) 10-50 mg/l, dose of
polymer ligand 1 g/l, shaking rate 120 rpm, rt.
288 Sayaka Fujita and Nobuo Sakairi
Complexation Stoichiometry
The ED-ch exhibited a high removal efficiency for Cu(II) by flocculation within a
narrow pH range, which is close to the IEP of ED-ch. It was also confirmed that most
Cu(II) could be removed at the pH value if excess amounts of EDTA residues were added
relative to Cu(II) (EDTA/Cu(II) molar ratio = 8.6). It is well known that EDTA forms
chelate complexes with Cu(II) in a 1:1 molar ratio in a range of pH values. As EDTA and
Cu(II) form 1:1 chelate complexes with an extremely high stability constant (log K =
18.7) [26], the excess amounts of ED-ch necessary for flocculation were unexpected. To
determine the stoichiometry between EDTA residues in ED-ch and Cu(II), we carried out
UV-vis titrations [27, 28] of Cu(II) with ED-ch at pH 5.5. As precipitate formation can
influence the absorbance of these samples, the experiment was carried out using ED-ch 6,
which is soluble over a wide range of pH.
Figure 7 shows the UV-vis spectra of ED-ch-Cu(II) chelates at pH 5.5. The spectra
show an adsorption maximum at 730 nm for the complex formed between Cu(II) and
different concentrations of ED-ch. The absorbance increased with increasing
concentrations of ED-ch. To determine the complex stoichiometry, the CEDTA/CCu was
plotted against absorbance of complex at 730 nm (Figure 7 right). Absorbance of the ED-
ch-Cu(II) complex increased in a nearly linear manner with increasing CEDTA/CCu, when
CEDTA/CCu was less than 1.0, and then reached plateau when CEDTA/CCu was more than 1.0.
This plateau of the plot (that is, absorbance was not changed with increasing CEDTA/CCu)
indicated that the chelation reaction had reached saturation. The intersection at the
turning point in the plot occurred at CEDTA/CCu = 1.0, suggesting that the EDTA residues
of ED-ch form chelation complex with Cu(II) in 1 : 1 molar ratio. This result was agreed
the stoichiometry of original EDTA with Cu(II).
Figure 7. UV-vis spectra of Cu(II) (7.87 10-4 mol/l) at pH 5.5 with ED-ch 6 (left) (from a to u: 0, 9.96
10-5, 1.99 10-4, 2.99 10-4, 4.00 10-3, 4.98 10-4, 5.97 10-4, 6.97 10-4, 7.97 10-4, 8.96 10-4,
9.96 10-4, 1.10 10-3, 1.19 10-3, 1.29 10-3, 1.39 10-3, 1.49 10-3, 1.59 10-3, 2.69 10-3, 1.79
10-3, 1.89 10-3, and 1.99 10-3 mol/l). Plot of absorbance at 730 nm against CEDTA/Ccu (right).
Chitosan Based Flocculants for the Removal of Heavy Metal Ions 289
As described in the previous section, the EDTA residues in ED-ch form chelate
complexes with Cu(II) in a 1:1 molar ratio at pH 5.5. Therefore, we expected that
discovery of new flocculation method other than pH control would increase remarkably
the removal efficiency of ED-ch. The removal ability under higher Cu(II) concentration
was investigated to clarify the effect of complexation stoichiometry on flocculation.
Experiments on the removal of Cu(II) by flocculation with ED-ch 4 were carried out by
changing the EDTA/Cu(II) molar ratio at pH 5.5, which is not closed to the IEP of ED-ch
4 (IEP = 3.0), and the results are shown in Figure 8. The theoretical removal efficiencies,
shown by dashed lines in the figure, were calculated by assuming 1:1 complexation
between the EDTA residues in ED-ch and Cu(II) and complete separation of ED-ch–
Cu(II) complexes from aqueous solution. The zeta potential of the mixture solution of
ED-ch and Cu(II) are also shown in figure.
Figure 8. Effect of EDTA/Cu(II) molar ratio on removal of Cu(II) by ED-ch 4 (left) and zeta potential
of mixture solution of ED-ch (right) at pH 5.5. Initial concentration of Cu(II): 10 mg/l, shaking rate:
120 rpm, rt.
The removal efficiency initially increased with increasing the molar ratio of
EDTA/Cu and the maximum removal was almost 100% at the molar ratio of 1.26-1.35.
Subsequently the removal efficiency gradually decreased with increasing the molar ratio.
The zeta potential increased with increasing the molar ratio of EDTA/Cu(II), and reached
approximately zero when the molar ratio was 1.2. This condition is the same as the
optimum conditions for high removal efficiency. These results indicated that the
flocculation process is controlled by the charge neutralization mechanism. Subsequently,
when the removal efficiency decreased with increasing molar ratios (above the optimum
molar ratio), the zeta potential also decreased to negative values, preventing flocculation
[29, 30, 31]. These charges should increase repulsion among particles, and thereby hinder
flocs growing into larger ones.
290 Sayaka Fujita and Nobuo Sakairi
We thought about factors that could affect charge neutralization to achieve effective
flocculation of ED-ch. EDTA is known to have four dissociation constants, and exists as
a zero-valent anion (H4Y) to a tetravalent anion (Y4-) at various pH values [32]. As EDTA
exists as a tetravalent anion (Y4-) under neutral and basic conditions, the EDTA residues
in ED-ch are expected to exist as trivalent anions, i.e., with three carboxylate ions. When
an EDTA residue forms a chelate complex with Cu(II), EDTA residue exist as
monovalent anion because two carboxylate ions are used by chelate formation. Therefore,
the coordination with Cu(II) should decrease the negative charges of the EDTA
residues in ED-ch.
CONCLUSION
A novel chitosan flocculant for removal of heavy metal ion was easily obtained by N-
acylation of chitosan with EDTA which is a strong chelating agent. Having property as
an amphoteric polyelectrolyte, EDTA grafted chitosan (ED-ch) had good water solubility
in both acidic and basic regions and precipitated in a narrow pH region, which is close to
its isoelectric point. The ED-ch could remove heavy metal ion from aqueous solutions,
and Cu(II) was almost completely removed by flocculation/precipitation process. The
removal of Cu(II) could be controlled by the pH of solutions as well as initial
concentration of the flocculant. Zeta-potential measurement of the Cu(II) solution in the
presence of ED-ch proved that the flocculation could be explained by the charge
neutralization. ED-ch can easily get rid of heavy metal ions from aqueous solution, and
are expected to be applicable for remediation and treatment system of wastewater.
REFERENCES
[1] Jian-Ping, W., Yong-Zhen, C., Yi, W., Shi-Jie, Y., Guo-Ping, S. & Han-Qing, Y.
(2012). A novel efficient cationic flocculant prepared through grafting two
monomers onto chitosan induced by Gamma radiation, RSC Adv., 2, 494–500.
[2] Hangcheng, Z., Yong, Z., Xiaogang, Y., Hongyi, L., Xiumei, Z. & Juming, Y.
(2015). An eco-friendly one-step synthesis of dicarboxyl cellulose for potential
application in flocculation, Ind. Eng. Chem. Res., 54, 2825–2829.
[3] Jaing-Ping, W., Shin-Jie, Y., Yi, W. & Han-Qing, Y. (2013). Synthesis,
characterization and application of a novel starch-based flocculant with high
flocculation and dewatering properties, Water Res., 47, 2643–2648.
Chitosan Based Flocculants for the Removal of Heavy Metal Ions 291
[4] Gregory, V. K., Hyun-Shik, C., Anatoly, I. F. & John, F. F. (2007). Structural study
of the incorporation of heavy metals into solid phase formed during the oxidation of
EDTA by permanganate at high pH, Environ. Sci. Technol., 41, 2560–2565.
[5] Fenglian, F. & Qi, W. (2011). Removal of heavy metal ions from wastewaters: A
review, J. Environ. Manage., 92, 407–418.
[6] Z. V. P. M. & Latesh, B. C. (2009). Separation of binary heavy metals from
aqueous solutions by nanofiltration and characterization of the membrane using
Spiegler–Kedem model, Chem. Eng. J., 150, 181–187.
[7] Terhi, S., Henrikki, L., Osmo, H. & Jouko, N. (2013). Coagulation–flocculation
treatment of municipal wastewater based on anionized nanocelluloses, Chem. Eng.
J., 231, 59–67.
[8] A. L. A., S. S. W., T. T. T. & A. Z. (2008). Improvement of alum and PACl
coagulation by polyacrylamides (PAMs) for the treatment of pulp and paper mill
wastewater, Chem. Eng. J., 137, 510–517.
[9] Chai, S. L., John, R. & Mei, F. C. (2014). A review on application of flocculants in
wastewater treatment, Process Saf. Environ. Prot., 92, 489–508.
[10] Zhen, Y., Hu, Y., Ziwen, J., Xin, H., Haibo, L., Aimin, L. & Rongshi, C. (2013). A
new method for calculation of flocculation kinetics combining Smoluchowski
model with fractal theory, Coll. Surf. A Physicochem. Eng. Asp., 423, 11-19.
[11] Alfredo, C., Rabindranath, B. M. & Russell, A. B. (1995). Synthesis and
characterization of cyclomaltoheptaose-based metal chelants as probes for intestinal
permeability, Carbohydr. Res., 267, 49-63.
[12] Naoji, K., Nobuhide, T., Takayuki, S. & Kaori, T. (2000). A simple preparation of
half N-acetylated chitosan highly soluble in water and aqueous organic solvents,
Carbohydr. Res., 324, 268-274.
[13] Katsutoshi, I., Keisuke, O., Kazuharu, Y., Tomoo, Y. & Takeshi, T. (1997).
Adsorption of Lead(II) Ion on Complexane Types of Chemically Modified
Chitosan, Bull. Chem. Soc. Jpn., 70, 2443-2447.
[14] Xiaochen, S., Leyang, Z., Xiqun, J., Yong, H. & Jian, G. (2007). Reversible Surface
Switching of Nanogel Triggered by External Stimuli, Angew. Chem. Int. Ed., 46,
7104-7107.
[15] Sayaka, F. & Nobuo, S. (2016). Water soluble EDTA-linked chitosan as a
zwitterionic flocculant for pH sensitive removal of Cu(II) ion, RSC Adv., 6, 10385–
10392.
[16] Lukas, G., Christopher, L., Wolfgang, H. G. & Aldo, R. B. (2018). Fabrication and
characterization of copper(II)-chitosan complexes as antibiotic-free antibacterial
biomaterial, Carbohydr. Polym., 179, 370-378.
[17] Jin, X., Stephen, P. M. & Richard, A. G. (1996). Chitosan Film Acylation and
Effects on Biodegradability, Macromolecules, 29, 3436-3440.
292 Sayaka Fujita and Nobuo Sakairi
[18] Joris, R. & Koen, B. (2014). Adsorption and chromatographic separation of rare
earths with EDTA- and DTPA-functionalized chitosan biopolymers, J. Mater.
Chem. A, 2, 1530-1540.
[19] Yong, R., Hayder, A. A., Fengbo, H., Hong, P. & Kaixun, H. (2013). Magnetic
EDTA-modified chitosan/SiO2/Fe3O4 adsorbent: Preparation, characterization, and
application in heavy metal adsorption, Chem. Eng. J., 226, 300-311.
[20] Can, Z., Qineng, P., Hongjuan, Z. & Jian, S. (2003). Synthesis and characterization
of water-soluble O-succinyl-chitosan, Eur. Polym. J., 39, 1629-1634.
[21] Varawut, T., Noppong, P. & Vipavee, P. H. (2003). Surface modification of
chitosan films.: Effects of hydrophobicity on protein adsorption, Carbohydr. Res.,
338, 937-942.
[22] Kai, Z., Johanna, H., Dieter, P., Margit, G., Thomas, G. & Steffen, F. (2010). NMR
and FT Raman characterisation of regioselectively sulfated chitosan regarding the
distribution of sulfate groups and the degree of substitution, Polymer, 51, 4698-
4705.
[23] Silvio, A., Roberto, G., Rita, N. & Enrica, S. (1987). 13C solid state CP/MAS NMR
studies of EDTA complexes, Inorganica. Chimica. Acta., 129, L23-25.
[24] Costas, S. P., Leo, R. S., Steven, P. A. & Norman, C. B. (1999). Precipitation of a
Water-Soluble ABC Triblock Methacrylic Polyampholyte: Effects of Time, pH,
Polymer Concentration, Salt Type and Concentration, and Presence of a Protein,
Langmuir, 15, 1613-1620.
[25] Eveliina, R., Jolanta, K. W., Tonni, A. K. & Mika, E. T. S. (2010). Adsorption of
Co(II) and Ni(II) by EDTA- and/or DTPA-modified chitosan: Kinetic and
equilibrium modeling, Chem. Eng. J., 161, 73-82.
[26] R. W. S. & C. N. R. (1956). A Rapid Electrochemical Method for the
Determination of Metal Chelate Stability Constants, J. Am. Chem. Soc., 78, 5513-
5518.
[27] Ying, D. & Leonard, F. L. (2003). A three-ring, linked cyclam derivative and its
interaction with selected transition and post-transition metal ions, Coord. Chem.
Rev., 245, 11–16.
[28] Lihua, L., Yanhong, L., Xing, L., Zhihua, Z. & Yulin, L. (2014). Chelating stability
of an amphoteric chelating polymer flocculant with Cu(II), Pb(II), Cd(II), and
Ni(II), Spectrochim. Acta., Part A, 118, 765–775.
[29] E. A. L., M. T. O., J. L. J. & A. O. (2014). Coagulation–flocculation mechanisms in
wastewater treatment plants through zeta potential measurements, J. Hazard.
Mater., 279, 1–10.
[30] Zhen, Y., Hu, Y., Ziwen, J., Tao, C., Haijiang, L., Habio, L., Aimin, L. & Rongshi,
C. (2013). Flocculation of both anionic and cationic dyes in aqueous solutions by
the amphoteric grafting flocculant carboxymethyl chitosan-graft-polyacrylamide, J.
Hazard. Mater., 254–255, 36–45.
Chitosan Based Flocculants for the Removal of Heavy Metal Ions 293
[31] Tingting, Z., Mingming, W., Wang, Y., Zhen, Y., Yuping, W. & Zhenggui, G.
(2014). Synergistic removal of copper(II) and tetracycline from water using an
environmentally friendly chitosan-based flocculant, Ind. Eng. Chem. Res., 53,
14913–14920.
[32] Yanfeng, L., Hui, P., Jinchuan, W., Jiaoxia, S. & Yuanliang, W. (2011). Novel
amphoteric pH-sensitive hydrogels derived from ethylenediaminetetraacetic
dianhydride, butanediamine and amino-terminated poly(ethylene glycol): Design,
synthesis and swelling behavior, Eur. Polym. J., 47, 40–47.
In: Flocculation: Processes and Applications ISBN: 978-1-53614-339-3
Editor: Eleonora Vollan © 2019 Nova Science Publishers, Inc.
Chapter 15
ABSTRACT
apply the Bratby method in the characterization of the turbidity removal process, through
the determination of the kinetic aggregation coefficient (KA) of the flocs and the kinetic
coefficient of rupture (KB) of the flocs. The experimental data were obtained by turbidity
determination (uT) of turbid water, consisting of a liquid solution containing kaolin and
humic acid. The coagulating agent used was iron chloride. The results were evaluated
based on flocculation kinetics by means of the aggregation constants (KA) and rupture
(KB). It was observed that floc aggregation resulted in the removal of 80% of turbidity,
with sedimentation time in 80 minutes and mean velocity gradient (GT) at 100 rpm,
resulting in a ratio of the kinetic coefficients of fracture and aggregation flock (KB/KA)
equals 0.005.
1. INTRODUCTION
According to Oliveira (2008) during flocculation, the agitation imposed on the liquid
medium promotes two effects simultaneously: aggregation and rupture (Figure 1).
Basically, the aggregation is the result of the encounters of the destabilized particles,
being that the agitation promotes a higher rate of encounters, forming agglomerates or
flocs; in the rupture occurs the erosion of the flocs by shearing forces, which are usually
accentuated with intense agitation or a great flocculation time, causing the partial or total
degradation of the flocs in a few seconds.
number of primary particles equal to the remaining turbidity, is similar to Equation (1)
given by :
𝑑𝑛
𝑑𝑡
= −𝐾𝐴 𝑛𝑇 𝐺𝐹 + 𝐾𝐵 𝑛0 (𝐺𝐹 )2 (1)
In which:
n0 = number of primary particles per unit volume at time t = 0;
nT = number of primary particles per unit volume at time t;
dn/dt = variation of number of primary particles per unit volume with according to
time;
KA = coefficient of aggregation;
KB = coefficient of rupture; and
GF = mean velocity gradient during flocculation.
𝑁0 𝐾 𝐾 −1
𝑁1
= [𝐾𝐵 𝐺𝐹 + (1 − 𝐾𝐵 𝐺𝐹 ) 𝑒 −𝐾𝐴 𝐺𝐹 𝑇𝐹 ] (2)
𝐴 𝐴
In which:
N0= initial turbidity of the supernatant;
N1=final turbidity of the supernatant; and
TF=flocculation time.
Rearranging Equation (2), we have Equation (3):
𝐾
1 (1− 𝐵𝐺𝐹 )
𝐾𝐴
𝐾𝐴 = 𝑙𝑛 [ ] (3)
𝐺𝐹 𝑇𝐹 1 𝐾
( 𝑁 − 𝐵𝐺𝐹 )
0 𝐾𝐴
𝑁
𝐾𝐵 1
= 𝑁 (4)
𝐾𝐴 𝐺𝐹 0
𝑁
2. METHODOLOGY
Preliminary tests were performed to evaluate the removal of the parameter turbidity
present in the water in the study, in which the composition consists of the presence of
humic acid at the initial concentration of 4 mg/L and caulin at a concentration of 5 g/L of
humic acid as the stock solution, then it was diluted to obtain solutions of turbidity values
150 uT, for the ferric chloride inorganic coagulating agent. To adjust the pH, 0.1 M
sodium hydroxide solution and 1M hydrochloric acid were used. It was used a 1 g/L
concentration solution for the coagulant ferric chloride.
Coagulation / flocculation / sedimentation assays were performed, aliquots being
removed at various times (1 to 60 minutes) and turbidity parameters were determined.
The turbidity was determined in a HACH DR / 2010 spectrophotometer at wavelength
860 nm, according to a procedure determined by Standard Methods (APHA, 2012).
The coagulation / flocculation assays were performed in Jar Test brand MILAN
Scientific Equipment. The pH parameter was determined in pHmeter TECNAL model
pHmeter TEC-2. The speeds used in the coagulation / flocculation assays to provide fast
mixing was set at 100 rpm and a fast mixing time of 1 min, while that the speed to
provide the slow mixing at 10 rpm and slow mixing time of 10 minutes, being the
sedimentation time = 60 minutes during the development of the experiments
(MADRONA, 2010).
The kinetic study of the aggregation and rupture of the particles by the Bratby
method was determined for maximum time of removal of 80 minutes (collections of 1
min, 20 min, 40 min, 60 min and 80 min), for fast mixing at 100 rpm and fast mixing
time of 1 minute, slow mixing at 10 rpm and slow mixing time of 10 minutes, settling
time = 80 minutes during the development of the experiments.
Figure 2 shows the flowchart of the kinetic study performed for the coagulation /
flocculation process.
Figure 2. Flowchart of the methodology used for the kinetic study of the coagulation / flocculation
process.
Determination of the Kinetic Coefficient of Aggregation … 299
3. RESULTS
The results for the kinetic study of aggregation and rupture (KB/KA) (Equation 4) in
the coagulation / flocculation / sedimentation process are shown in Table 1, which
indicates the turbidity parameter removal data obtained in the experiment and turbidity
initial dose is 150 uT.
Figure 3 represents a graphic of N0/N (Equation 4) which stands for the ratio of initial
turbidity to final turbidity versus time.
After the linear adjustment (Figure 3), it was verified that the ratio of turbidity
parameter removal increased as a function of time in the coagulation / flocculation
process with the use of the chemical coagulant ferric chloride.
Figure 4 shows the graphic of the coefficient of rupture with the coefficient of
aggregation (KB/KA) (Equation 4) as a function of time. According to Moreira (2003) the
size of the flocs, once formed or aggregated, any disturbance of the system, however
small, may cause its rupture.
300 J. D. Peruço Theodoro, P. S. Theodoro and R. Bergamasco
Figure 4. Relationship of the coefficient of rupture and aggregation (KB/KA) as a function of time.
Thus, according to Figure 4, the KB/KA ratio is inversely proportional to time. Figure
5 shows a graphic of KB/KA as a function of turbidity parameter.
Figure 6. Relation of initial turbidity and final turbidity (N0/N) as a function of the ratio of the
coefficient of rupture and coefficient of aggregation (KB/KA).
Determination of the Kinetic Coefficient of Aggregation … 301
In Figure 6, the removal ratio of the turbidity parameter was found to decrease with
the ratio of the coefficient of rupture to the coefficient of aggregation in the water
treatment process with the use of ferric chloride as a coagulant.
According to Moreira (2003) that the chemical coagulant aluminum sulphate
Al2(SO4)3.18H2O was used for the pH of coagulation / flocculation to values between 6.0
and 6.5 with reason to allow the prevailing coagulation mechanism by scanning, it was
concluded that the higher the coefficient of aggregation tended to be, the lower the
relationship between the flocculation times in the batch reactor and the continuous flow
reactor.
Many water treatment plants (ETAs) that work in hydraulic overload generally
choose to work with iron salts in preference to chemical coagulants based on aluminum
salts. This is happening, because the iron allows to obtain higher values of aggregation
constants and lower values of rupture constants when compared with aluminum salts, the
quality of the decanted water can be significantly improved for a flocculation system
operating with the same time of hydraulic detention (MOREIRA, 2003).
Bache and Gregory (2010) affirm that the size of the floc can be pointed as the factor
that exerts greater influence on the density and the force of formation of flocs. According
to Bache et al., (1997) and Jarvis et al., (2005) the strength of the floc is directly related
to the structure of the floc and, therefore, dependent on the formation process of the flocs.
The flocs do not grow infinitely. The growth reaches a steady state for a given shear
condition. Generally the growth of the flocs is related to the break, so the rate of
aggregation is the balance between the formation and rupture of the aggregates.
CONCLUSION
REFERENCES
APHA – American Public Health Association. Standard Methods for the Examination of
Water and Wastewater. 22 ed. Washington, 2012.
Argaman, Y. and Kaufman, W. J. (1970) Turbulence and Flocculation. J. Sanit. Eng. Div.
ASCE 96, 223.
302 J. D. Peruço Theodoro, P. S. Theodoro and R. Bergamasco
Bache, D. H.; Johnson, C.; Mcgillian, J. F.; Rasool, E. A conceptual view of floc
structure in the sweep floc domain. Water Science and Technology. v.36, p. 49-56,
1997.
Bache, D. H. and Gregory, R. Flocs and separation processes in drinking water treatment:
a review. Journal of Water Supply: Researche and Technology-AQUA. v.59, n.1,
2010.
Bratby, J. et al. Design of flocculation systems from batch test data. Water SA. v.3, n. 4,
p. 173-182, 1977.
Di Bernardo, L. et al. Use of mathematical modeling for the design of mechanized
flocculation chambers in series in water treatment plants. Sanitary and
Environmental Engineering, v. 10, n. 1, p. 82-90, 2005.
Filho, S. S. F., Physical-Chemical Processes I - Coagulation. Escola Politécnica da USP
Department of Hydraulic and Sanitary Engineering- (2010).
Madrona G. S. Extraction / Purification of the Moringa oleifera Lam Seed Active
Compound and its use in the Treatment of Water for Human Consumption. Thesis of
Doctorate in Chemical Engineering, UEM, Maringá -PR, 2010.
Moreira, H. A.- The Flocculation Process in the Treatment of Water Supply:
Transposition of Jury Test Results for Real-Scale Flocculation Systems -22º
Brazilian Congress of Sanitary and Environmental Engineering Joinville - Santa
Catarina - I-082, 14-19 of September 2003.
Oliveira, D. S. Evaluation of the turbidity removal efficiency due to variations in the
length of helical tubular flocculators / Danieli Soares de Oliveira. Dissertation
(master’s degree) - Federal University of Espírito Santo, Technological Center p.
116, 2008.
Jarvis, P.; Jefferson, B.; Gregory, J. and Parsons, S.A. A Review of floc strength and
breakage. Water Research. v.39, p. 3121-3137, 2005.
INDEX
algal cells, 116, 117, 123, 124, 126, 131, 132, 212,
A
214
algal organic matter, vii, x, 107, 108, 110, 111, 131,
acetic acid, 25, 108, 280, 281, 282
134, 135, 137, 139, 140, 141, 142, 216
acid, viii, xii, xv, 13, 14, 15, 16, 18, 20, 21, 22, 23,
alkalinity, 172, 262, 264, 266, 267, 275
25, 26, 27, 36, 37, 39, 43, 48, 50, 72, 74, 76, 77,
aluminium, vii, x, 5, 6, 32, 116, 118, 120, 121, 122,
108, 111, 122, 134, 136, 158, 187, 190, 200, 201,
124, 134, 137, 143, 264, 265, 267, 270, 271
219, 229, 258, 265, 267, 269, 272, 277, 278, 279,
aluminum sulfate, vii, ix, 47, 48, 50, 56, 69, 145
280, 281, 282, 284, 296, 298
amino acids, 110, 112, 113, 120, 126, 141, 268
acidic, x, xv, 8, 22, 29, 30, 72, 107, 115, 116, 117,
ammonium ions, xiv, 218, 225, 229, 236, 237, 245,
120, 122, 125, 126, 158, 203, 208, 226, 229, 264,
251, 253
267, 268, 270, 277, 278, 285, 290
amorphous precipitate, 249
activated carbon, 15, 109, 131, 134, 141, 178, 214,
anaerobic digestion, 66
215, 235, 256
aquatic systems, 203, 278
acylation, viii, xv, 277, 279, 282, 290
aqueous solutions, 4, 38, 41, 43, 113, 285, 290, 291,
adsorption, 3, 6, 7, 21, 26, 27, 38, 83, 84, 85, 86, 87,
292
88, 89, 109, 118, 120, 121, 122, 125, 131, 134,
aromatic compounds, 4, 263
136, 137, 140, 141, 157, 170, 177, 178, 182, 208,
aromatic rings, 35
209, 215, 221, 235, 242, 243, 249, 251, 254, 255,
assisted sedimentation, viii, xiv, 259, 268
256, 257, 261, 265, 268, 278, 283, 288, 291, 292
atmosphere, 4, 54, 99
advanced oxidation process, 170, 176, 270
agglomeration, 2, 50, 83, 85, 220, 223
aggregation, vi, viii, 2, 82, 83, 84, 85, 86, 87, 88, 89, B
97, 98, 103, 105, 121, 138, 220, 260, 266, 278,
295, 296, 297, 298, 299, 300, 301 bacteria, 8, 9, 10, 11, 17, 18, 20, 22, 27, 50, 56, 82,
aggregation process, 84, 87, 138 156
agriculture, 30, 48, 66, 102, 171 bacterium, 11, 13, 20, 21, 28, 34
aldehydes, 110, 112, 113 bicarbonate, 229, 264
algae, viii, ix, 1, 10, 18, 22, 23, 81, 82, 108, 109, biochemistry, 136
111, 116, 117, 122, 124, 126, 130, 131, 135, 136, biocompatibility, 3, 10, 23
137, 138, 139, 156, 198, 200, 208, 210, 214, 215, bioconversion, 136
216, 223, 235, 266 biodegradability, viii, 1, 2, 3, 4, 8, 10, 17, 30, 35,
172, 219, 222, 229, 234, 235, 251, 273
304 Index
biodegradation, 38, 219, 229, 236, 275 chelating, xv, 27, 170, 178, 277, 278, 290, 292
biofuel, 116, 126 chemical, vii, viii, ix, x, xi, 1, 2, 3, 4, 7, 8, 10, 11, 12,
biogas, 66, 99, 101, 102, 219 13, 14, 15, 16, 21, 22, 23, 24, 26, 28, 29, 30, 32,
biological activity, 11, 43 33, 42, 43, 47, 48, 66, 69, 81, 83, 85, 87, 97, 102,
biological processes, 99 104, 116, 143, 144, 145, 147, 150, 152, 155, 156,
bioluminescence, 40 157, 166, 169, 170, 171, 174, 176, 180, 181, 186,
biomass, 11, 21, 24, 32, 34, 36, 49, 102, 103, 116, 198, 199, 215, 222, 223, 225, 226, 229, 256, 260,
123, 126, 136, 235, 251, 278 262, 264, 265, 269, 270, 272, 274, 275, 278, 295,
biomaterials, 2, 40 299, 301
biomedical applications, 37 chemical bonds, 260
biomolecules, 6, 45 chemical characteristics, ix, 69
biopolyelectrolytes, vi, xi, 169, 170, 171, 172, 173, chemical interaction, 85
174, 176 chemical properties, 3, 33
biopolymers, ix, 2, 4, 8, 10, 22, 24, 25, 26, 32, 35, chemical reactions, ix, 47, 48
38, 41, 43, 47, 48, 49, 51, 52, 56, 57, 58, 59, 61, chemical reactivity, 3
62, 64, 67, 86, 113, 122, 124, 131, 223, 226, 243, chitosan, vi, viii, xii, xiv, xv, 2, 6, 8, 12, 19, 23, 24,
292 25, 27, 28, 30, 31, 33, 35, 37, 38, 39, 42, 43, 44,
bioremediation, 22, 40 116, 117, 124, 125, 129, 135, 169, 171, 172, 173,
biosafety, viii, 1, 2, 30, 31 174, 175, 176, 178, 209, 210, 211, 213, 214, 215,
biosensors, 41 218, 222, 226, 227, 239, 241, 242, 243, 244, 245,
biosynthesis, 10, 44, 137 246, 249, 250, 251, 252, 253, 254, 255, 258, 267,
biotechnological applications, 12, 20, 31 269, 277, 278, 279, 280, 281, 282, 283, 284, 285,
biotechnology, 23, 65, 177 286, 287, 290, 291, 292, 293
blue-green algae, 198, 200 chlorination, 108, 133, 136, 166, 215, 235
Bratby method, viii, 296, 298, 301 chlorine, 48, 82, 138, 166, 275
Brazil, vii, viii, ix, xii, 1, 44, 69, 70, 71, 76, 78, 79, chromatography, 122, 272
81, 91, 92, 143, 145, 152, 155, 157, 165, 166, chromium, 27, 144, 152, 177, 238
179, 180, 186, 187, 188, 295 climate, 54, 99, 223
coagulant, v, vii, viii, ix, xi, xii, xiii, xiv, 7, 13, 16,
31, 35, 38, 41, 48, 49, 51, 53, 54, 57, 58, 59, 60,
C
61, 62, 64, 65, 69, 70, 72, 74, 75, 78, 86, 116,
117, 119, 121, 123, 124, 125, 126, 128, 129, 130,
carbon, 11, 15, 28, 34, 43, 44, 49, 109, 131, 132,
131, 133, 136, 143, 144, 145, 146, 147, 149, 150,
134, 136, 141, 170, 171, 172, 177, 178, 200, 214,
152,155, 157, 158, 159, 161, 163, 164, 165, 166,
215, 226, 235, 256, 262, 284
167, 171, 172, 174, 176, 178, 179, 181, 185, 193,
carbon tetrachloride, 43
195, 197, 199, 201, 202, 207, 208, 209, 210, 211,
carboxyl, 6, 14, 21, 26, 126, 282, 285
212, 215, 218, 220, 222, 223, 226, 239, 242, 244,
carboxylic acids, 112, 268
245, 246, 249, 250, 251, 252, 253, 255, 257, 258,
carboxylic groups, 27
261, 262, 263, 264, 265, 266, 267, 268, 269, 270,
carboxymethyl cellulose, 26, 32
271, 274, 275, 298, 299, 301
cell surface, 116, 117, 122, 127
coagulation mechanism, 108, 117, 118, 136, 176,
cellular organic matter, 109, 113, 128, 132, 138, 139,
215, 301
141
coagulation process, 8, 70, 180, 209, 213, 265, 295
cellulose, 2, 3, 6, 10, 11, 12, 15, 17, 23, 26, 29, 32,
coagulation-flocculation, vii, viii, xi, xiii, 15, 16, 25,
33, 35, 37, 38, 40, 43, 44, 177, 270, 273, 275, 290
29, 33, 36, 42, 64, 71, 73, 74, 76, 77, 78, 145,
charge density, 7, 8, 87, 115, 124, 171, 172, 173,
149, 155, 169, 170, 171, 172, 173, 174, 176, 186,
174, 176, 223, 243
197, 198, 199, 201, 202, 203, 204, 206, 207, 208,
charge neutralization, xv, 5, 6, 7, 26, 28, 108, 118,
209, 212, 213, 215, 235, 256, 263, 270, 271, 273,
122, 123, 176, 203, 208, 261, 265, 277, 289, 290
274
Index 305
colloids, 6, 8, 29, 34, 37, 38, 83, 84, 96, 133, 141, dosage, xiii, 6, 8, 9, 22, 24, 30, 139, 174, 197, 199,
215, 220, 221, 222, 242, 260, 263, 265, 268, 278 201, 202, 203, 204, 205, 206, 207, 209, 210, 212,
composition, vii, viii, x, xii, 8, 16, 20, 21, 22, 26, 30, 216, 253, 262, 265, 267
39, 49, 53, 89, 107, 110, 128, 129, 130, 135, 136, drinking water, x, xiii, 48, 49, 62, 63, 64, 66, 67, 79,
137, 144, 179, 219, 224, 238, 261, 298 107, 108, 110, 126, 133, 134, 138, 140, 166, 197,
compounds, vii, viii, x, xiii, 4, 29, 44, 49, 82, 107, 198, 199, 200, 208, 213, 214, 215, 216, 260, 266,
108, 109, 110, 112, 115, 118, 120, 122, 124, 125, 267, 273, 295, 302
126, 131, 136, 197, 198, 219, 226, 229, 236, 237,
242, 249, 250, 251, 252, 253, 260, 263, 268, 269,
E
270, 272, 275, 278
compression, 28, 84
effluent, viii, x, xii, xiv, 11, 17, 20, 28, 29, 30, 32,
conductivity, x, xii, xiii, 143, 145, 149, 150, 171,
33, 38, 41, 44, 92, 95, 97, 98, 99, 101, 118, 143,
172, 180, 181, 182, 187, 188, 190, 193, 194, 225,
144, 145, 146, 147, 149, 150, 152, 153, 167, 170,
231, 234, 235, 262
179, 180, 181, 182, 185, 186, 188, 191, 259, 260,
consumption, xiii, 53, 64, 70, 95, 101, 144, 170, 174,
262, 265, 266, 267, 273, 274, 275
186, 187, 190, 191, 193, 194, 217, 218, 266, 267
electric conductivity, xii, 150, 181, 187
consumption patterns, xiii, 217, 218
electrical conductivity, x, xii, 143, 149, 150, 171,
contaminated water, 278, 295
172, 179, 181, 182, 188, 193, 194
contamination, 2, 27, 53, 170, 219, 223
electrocoagulation, vi, viii, xii, xiv, 170, 177, 187,
coordination, 27, 140, 290
188, 190, 191, 193, 194, 195, 259, 269
copolymer, 16, 26, 28, 37, 38
electrostatic interaction, 6, 26, 27, 85, 117, 120, 121,
copolymerization, 12, 13, 23
125, 126, 127, 171, 174, 254
cost, viii, ix, xii, 1, 2, 5, 10, 12, 13, 14, 15, 16, 17,
energy, ix, xiii, 12, 47, 66, 82, 84, 85, 87, 88, 89, 90,
21, 25, 30, 34, 47, 70, 99, 102, 109, 131, 177,
91, 93, 94, 96, 97, 99, 101, 102, 103, 130, 134,
187, 188, 191, 194, 199, 222, 242, 268, 278
177, 187, 190, 191, 194, 219, 261, 262, 263, 271,
cost effectiveness, 199
274
cotton, 74, 159, 160, 274, 282
energy consumption, xiii, 101, 187, 190, 194
covalent bond, 121
environmental conditions, 4, 110
cultivation, 22, 110, 123, 136
environmental impact, viii, x, xii, xiii, 48, 49, 50,
culture, 8, 11, 20, 28, 66, 111, 114, 116, 136
143, 144, 179, 217
culture media, 8
environmental issues, 236
culture medium, 20, 28, 116
erosion, 82, 262, 296, 297
cyanobacteria, xiii, 18, 27, 64, 109, 110, 111, 133,
ethylenediaminetetraacetic acid, 279
135, 137, 138, 197, 198, 208, 209, 210, 211, 212,
exopolysaccharides, 20, 22, 32, 35, 137
213, 214, 215, 216
experimental design, 191, 222, 244, 249
extracellular microcystins, 199, 201, 202, 212, 215
D extracellular organic matter, 109, 123, 135, 138
extraction, xi, 3, 7, 14, 30, 32, 33, 36, 37, 39, 43, 49,
degradation, 4, 11, 56, 70, 124, 229, 235, 243, 296 53, 54, 62, 70, 71, 72, 152, 159, 166, 169, 225,
denitrification, 99, 236, 237, 256 260, 274
derivatives, 10, 11, 12, 21, 28, 34, 42, 86, 171, 235 extracts, vii, ix, 47, 49, 57, 62, 64
detergents, xiv, 218, 235, 251, 253, 254
diatoms, 109, 111, 135, 136
F
disinfection, 82, 108, 134, 136, 198, 270, 275
dispersion, 12, 58, 95, 96, 97, 171, 174, 263, 266,
fats and oil, xii, 169, 170, 171, 172, 173, 174, 175
267
fermentation, viii, xiii, 1, 2, 6, 21, 30, 180, 217, 229
ferric chloride, vii, viii, ix, x, xiii, xiv, 5, 47, 48, 50,
62, 116, 117, 133, 143, 144, 145, 149, 171, 197,
306 Index
199, 200, 201, 202, 203, 204, 205, 206, 208, 211,
H
212, 213, 215, 218, 219, 221, 239, 241, 242, 243,
244, 245, 246, 249, 250, 252, 253, 264, 265, 267,
health, xi, xiii, 2, 5, 27, 32, 39, 70, 99, 103, 133, 156,
269, 270, 274, 296, 298, 299, 301
157, 160, 161, 163, 165, 171, 197, 198, 213, 214,
fertilizers, 29, 102, 235
215, 219, 222, 237, 251, 266
filters, xi, 73, 74, 155, 159, 198
health effects, 215
filtration, vii, ix, xi, xii, xiii, 7, 69, 70, 71, 73, 74, 75,
health problems, 2, 27
76, 77, 78, 81, 86, 109, 122, 128, 131, 133, 137,
health risks, 70
140, 155, 157, 159, 161, 162, 163, 164, 165, 170,
heavy metals, 15, 20, 22, 27, 35, 40, 156, 170, 176,
179, 180, 181, 182, 184, 185, 208, 217, 254, 269,
226, 238, 268, 270, 278, 291
280, 282, 295
human, xiii, 5, 27, 33, 52, 53, 64, 70, 71, 171, 186,
floc properties, x, 13, 108, 128, 129, 131, 134, 141
197, 198, 214, 219, 222, 237, 251, 278
flocculant, v, vii, viii, ix, xiv, xv, 1, 2, 6, 7, 8, 9, 11,
human health, xiii, 5, 27, 171, 197, 198, 219, 222,
13, 15, 16, 17, 18, 20, 21, 22, 23, 24, 25, 26, 27,
237, 251
28, 29, 30, 33, 34, 35, 37, 38, 39, 40, 41, 42, 43,
humic substances, 120, 122, 123, 126, 127, 131, 137,
44, 57, 58, 62, 65, 104, 172, 174, 176, 178, 214,
139, 142, 229, 242
216, 218, 220, 222, 226, 227, 239, 241, 242, 243,
hydrogen, 6, 7, 10, 11, 22, 86, 117, 120, 121, 122,
244, 245, 246, 249, 250, 251, 253, 254, 258, 262,
125, 127, 130, 229, 268, 272, 285
263, 265, 267, 268, 269, 274, 275, 277, 278, 279,
hydrogen bond(s), 6, 7, 10, 11, 22, 86, 117, 120, 121,
280, 284, 290, 291, 292, 293
122, 125, 127, 130, 285
flotation, ix, 81, 87, 98, 104, 109, 128, 134, 135,
hydrogen peroxide, 272
139, 214, 220, 226, 253, 263, 271, 273, 295
hydrolysis, 32, 84, 87, 118, 120, 203, 220, 221, 264,
food, viii, xi, 1, 2, 3, 6, 11, 14, 29, 34, 36, 37, 38, 39,
265, 266, 282
40, 41, 44, 45, 56, 66, 169, 171, 278
hydrophilicity, 3, 15, 111, 212
food chain, 278
hydrophobicity, 86, 111, 131, 235, 236, 262, 292
food industry, 14, 29, 44, 66, 171
hydroxide, 48, 72, 116, 118, 121, 158, 190, 201, 264,
formation, xiv, 2, 5, 9, 23, 25, 26, 27, 30, 39, 82, 83,
270, 298
85, 86, 87, 88, 95, 96, 108, 118, 122, 126, 133,
hydroxyl, 3, 6, 11, 12, 21, 26, 27, 120, 239, 264
134, 136, 138, 140, 198, 218, 223, 229, 237, 243,
hydroxyl groups, 3, 11, 12, 21, 26, 120
244, 260, 261, 263, 264, 265, 266, 269, 272, 283,
286, 288, 290, 295, 301
fouling, 108, 122, 135, 136, 139, 176, 213, 216, 262 I
freshwater, 18, 122, 139, 200
fungi, viii, 1, 9, 10, 17, 18, 21, 23, 53, 56, 57, 226 impurities, x, 57, 70, 95, 107, 128, 129, 131, 132,
156, 157, 260, 295
inorganic coagulants, xiv, 122, 144, 180, 218, 222,
G 266, 269
intermolecular interactions, 4, 130, 131
genus, 51, 64, 236
intracellular, xiii, 109, 136, 197, 199, 201, 202, 203,
glucose, 10, 20, 23, 111
204, 205, 206, 207, 208, 209, 210, 211, 212, 213
glycoproteins, 17, 20
ions, xiv, 4, 12, 15, 20, 22, 24, 27, 28, 38, 41, 42, 83,
glycosaminoglycans, 24
84, 125, 134, 178, 203, 218, 225, 229, 235, 236,
gravity, 49, 96, 157, 171, 261
237, 245, 251, 253, 261, 264, 265, 267, 268, 270,
green alga, 109, 111, 117, 126, 135, 138, 198, 200
273, 278, 290, 291
greenhouse gas emissions, 99
iron, vii, ix, 27, 47, 48, 49, 50, 70, 73, 82, 84, 137,
guar gum, 29, 39, 86, 173, 267
156, 159, 203, 221, 222, 237, 250, 256, 263, 264,
270, 296, 301
Index 307
neutral, xi, xii, xiii, 3, 7, 16, 22, 24, 26, 59, 72, 74,
K
112, 116, 120, 125, 155, 162, 180, 187, 193, 203,
208, 222, 226, 242, 285, 290
kinetic study, 296, 298, 299
neutralization, xv, 5, 6, 7, 26, 28, 83, 85, 108, 118,
kinetics, vii, x, 6, 34, 81, 86, 88, 265, 291, 296, 301
122, 157, 174, 176, 193, 203, 208, 221, 259, 260,
261, 262, 265, 269, 277, 289, 290
L nitrates, 218, 225, 236, 237, 245, 251, 252, 253
nitrification, 99, 100, 236, 237, 256
landfill leachate, 16, 31, 219, 222, 242, 245, 249, nitrogen, 29, 43, 44, 49, 97, 99, 100, 101, 102, 103,
250, 253, 254, 255, 256, 257, 258 104, 171, 172, 180, 226, 236, 237, 238, 249, 251,
lignin, 10, 24, 26, 28, 33, 37, 173 252, 253, 255, 262
nitrogen compounds, 44, 251, 252, 253
M nitrogen dioxide, 43
nitrogen gas, 236, 237
meat processing wastewater, vi, 169, 172, 173 nucleic acid, 18, 20, 21, 110, 120, 126, 134
mesquite gum, viii, xii, 169, 171, 173, 176 nutrients, xiii, 22, 180, 197, 198, 257
metabolites, 20, 133, 137, 199, 208, 212
metal ion, viii, xv, 12, 20, 27, 35, 38, 41, 42, 261, O
277, 278, 286, 290, 291, 292
metal salts, 28, 70, 134, 141, 214, 221, 222 oil, xii, 5, 11, 18, 20, 21, 29, 30, 32, 36, 42, 85, 86,
metals, 4, 5, 15, 20, 21, 22, 27, 35, 40, 48, 49, 50, 62, 170, 174, 175, 260, 272, 273
156, 170, 176, 221, 223, 226, 238, 246, 262, 265, oil production, 260
268, 270, 278, 291 organic coagulant(s), vii, ix, xi, 69, 74, 144, 147,
methodology, 4, 39, 53, 72, 145, 160, 190, 222, 244, 149, 155, 157, 158, 164, 165, 185, 195, 221, 222,
249, 255, 256, 274, 298 266, 267
methylene blue, 26, 97, 178 organic compounds, x, 49, 107, 108, 109, 110, 120,
microcystin, xiii, 133, 198, 199, 202, 203, 204, 205, 126, 131, 136, 229, 242, 249, 263
206, 208, 209, 210, 212, 213, 214, 215, 216 organic matter, vii, viii, x, xii, 33, 43, 48, 49, 57, 82,
microcystis aeruginosa, 198 100, 101, 107, 108, 109, 110, 111, 112, 113, 118,
microorganisms, ix, 2, 9, 10, 11, 17, 30, 49, 50, 59, 120, 121, 122, 123, 128, 131, 132, 133, 134, 135,
66, 81, 82, 101, 132, 156, 226, 229, 236, 243, 136, 137, 138, 139, 140, 141, 142, 157, 170, 179,
260, 268, 275 184, 203, 208, 214, 216, 219, 223, 229, 234, 236,
molecular weight, x, 4, 5, 7, 8, 20, 22, 24, 86, 87, 96, 242, 258, 260, 261, 262, 263, 266, 267, 270, 273,
107, 111, 113, 131, 157, 226, 243, 273, 279, 280 275
molecules, 3, 12, 24, 25, 30, 84, 85, 87, 88, 108, 113, organic polymers, 33, 243
120, 226, 236, 264, 268 organic solvents, 11, 291
Moringa oleifera, ix, xi, xii, 32, 47, 51, 64, 125, 155, oxidation, viii, xiii, xv, 12, 110, 124, 135, 136, 137,
156, 157, 158, 160, 161, 163, 164, 165, 166, 179, 139, 142, 170, 176, 177, 188, 191, 216, 217, 236,
181, 222, 256, 267, 302 237, 259, 263, 268, 270, 271, 272, 274, 275, 278,
291
N
P
nanocomposites, 38, 40, 274
nanomaterials, 38, 170 pectin, 10, 14, 15, 27, 32, 33, 34, 35, 37, 39, 40, 42,
nanoparticles, 21, 37, 274 43, 44
natural polymers, 6, 11, 31, 32, 117, 125 peptides, x, 107, 110, 111, 113, 114, 115, 118, 119,
natural resources, ix, 47, 144 120, 121, 122, 123, 126, 131, 134, 138, 141, 142
natural sciences, 64
308 Index
pH, ix, x, xi, xii, xiii, xiv, xv, 3, 5, 6, 7, 8, 13, 16, 20, polysaccharide(s), v, vii, viii, 1, 2, 3, 4, 6, 7, 8, 9, 10,
22, 23, 27, 35, 48, 58, 59, 60, 61, 62, 69, 71, 72, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23,
73, 74, 75, 76, 77, 78, 107, 115, 116, 117, 118, 24, 25, 26, 27, 28, 29, 30, 33, 34, 36, 37, 38, 39,
119, 120, 121, 122, 123, 124, 126, 128, 130, 131, 40, 41, 42, 45, 110, 111, 113, 115, 117, 120, 122,
132, 135, 139, 143, 145, 146, 147, 155, 158, 159, 124, 126, 129, 133, 136, 137, 278
160, 161, 162, 163, 164, 165, 169, 172, 173, 174, potassium, 136, 137, 142, 215, 265, 279
175, 179, 181, 182, 183, 187, 188, 189, 190, 191, potato, 12, 14, 22, 28, 29, 34, 40, 43
192, 193, 194, 197, 199, 200, 201, 202, 203, 204, potato starch, 22, 28, 29, 34, 40
205, 206, 207, 208, 210, 211, 212, 213, 218, 220, precipitation, xv, 7, 9, 116, 117, 118, 170, 220, 221,
221, 222, 223, 225, 226, 227, 229, 230, 231, 234, 265, 269, 278, 287, 290
242, 244, 246, 249, 250, 253, 261, 262, 263, 264, preparation, iv, vii, viii, xv, 1, 23, 35, 49, 57, 58,
267, 268, 269, 270, 277, 280, 281, 282, 285, 286, 158, 181, 277, 280, 282, 291
287, 288, 289, 290, 291, 292, 293, 298, 301 proteins, x, 12, 17, 20, 86, 107, 110, 111, 113, 114,
pharmaceutical, 3, 6, 11, 15, 17, 23, 45, 235 115, 118, 119, 120, 121, 122, 123, 126, 127, 129,
pharmacology, 2, 17 131, 134, 137, 138, 141, 142, 260, 268
phenol, xiv, 218, 219, 225, 232, 235, 238, 239, 241, pulp, 11, 14, 28, 32, 36, 38, 41, 42, 43, 72, 291
242, 245, 251, 253 purification, xiv, 3, 10, 30, 48, 64, 138, 188, 218,
phenolic compounds, 226, 260, 268, 269, 270, 272 273, 274, 279
phosphorus, xiv, 29, 30, 49, 101, 172, 180, 218, 225,
226, 230, 233, 235, 238, 245, 249, 250, 253
Q
photocatalysis, 170, 177, 272, 274
physical properties, 22, 134, 214
quality control, 105
physicochemical characteristics, 11, 175, 226, 238,
quaternary ammonium, 13
242
physicochemical properties, 3, 4, 8, 14
phytoplankton, x, 107, 109, 110, 112, 113, 114, 122, R
136, 139
plants, v, vii, viii, ix, x, xiii, 1, 2, 3, 7, 10, 11, 12, 14, raw materials, viii, 1, 5, 11, 56
16, 17, 30, 47, 48, 49, 50, 51, 52, 53, 54, 55, 58, reactants, 5, 88, 95
59, 62, 63, 65, 66, 67, 70, 81, 82, 89, 91, 94, 95, reaction temperature, 5
97, 98, 99, 101, 103, 107, 108, 109, 174, 178, reaction time, 5
197, 198, 235, 260, 265, 270, 292, 301, 302 reagents, xiv, 48, 85, 88, 218, 245, 248, 249, 250,
pollutant, xiv, 25, 29, 144, 188, 218, 223, 242, 286 251, 252, 253
pollution, viii, xii, 5, 27, 35, 99, 103, 144, 187, 219, recovery, 2, 25, 28, 29, 38, 66, 99, 101, 102, 171,
241, 242 219
polyacrylamide, 5, 11, 13, 15, 16, 23, 26, 29, 32, 37, removal of heavy metal ion, viii, xv, 277, 278, 290,
38, 41, 44, 226, 227, 257, 266, 273, 274, 278, 292 291
polyamides, 86 repulsion, 7, 9, 49, 82, 84, 208, 220, 242, 286, 289
polyelectrolyte complex, 23, 39
polyhydroxyalkanoates, 31 S
polymer, vii, ix, xiv, 3, 4, 6, 7, 8, 12, 23, 24, 25, 26,
33, 40, 69, 71, 72, 79, 86, 87, 88, 89, 117, 121, salts, 5, 28, 50, 70, 82, 84, 89, 116, 117, 129, 134,
124, 131, 171, 218, 222, 223, 242, 243, 253, 265, 135, 141, 182, 188, 194, 214, 221, 222, 243, 249,
266, 273, 274, 275, 286, 287, 292 265, 266, 275, 301
polymeric materials, 125, 176, 266 science, 33, 45, 104, 138, 214
polymers, viii, xiv, 1, 3, 6, 11, 25, 26, 31, 32, 33, 63, sedimentation, vii, viii, ix, x, xi, xii, xiv, 16, 18, 26,
70, 83, 86, 87, 89, 96, 104, 113, 116, 117, 121, 32, 47, 49, 50, 66, 69, 70, 71, 73, 74, 75, 76, 77,
124, 125, 170, 218, 223, 243, 258, 266, 270, 273, 78, 86, 87, 89, 109, 126, 128, 139, 140, 143, 145,
278 146, 149, 155, 157, 161, 162, 163, 164, 171, 179,
Index 309
181, 185, 208, 216, 259, 260, 261, 262, 268, 269, stoichiometry, xv, 277, 278, 281, 288, 289
270, 272, 275, 295, 297, 298, 299 sulfate, vii, ix, 5, 27, 44, 47, 48, 50, 56, 62, 69, 84,
seed, 16, 28, 32, 41, 157, 158, 163, 165, 166, 181, 138, 145, 200, 250, 264, 267, 270, 273, 274, 275,
249, 267, 274 278, 292
separation, ix, x, xi, 10, 35, 38, 81, 82, 86, 87, 96, synthetic polymers, 3, 86, 258
97, 102, 103, 104, 108, 109, 128, 133, 135, 136,
138, 139, 140, 141, 157, 169, 171, 173, 174, 178,
T
213, 215, 216, 226, 257, 268, 269, 278, 289, 291,
292, 302
tannin, xii, 144, 145, 146, 147, 149, 150, 151, 152,
sewage, 16, 17, 20, 21, 38, 44, 50, 63, 97, 100, 101,
179, 180, 181, 184, 185, 267
176
technology, 4, 11, 25, 104, 108, 130, 138, 166, 170,
shear, xiv, 83, 84, 85, 128, 129, 130, 131, 134, 141,
214, 215, 278
199, 208, 212, 218, 223, 261, 262, 263, 264, 275,
temperature, xiv, 3, 5, 7, 8, 9, 20, 48, 55, 56, 99, 110,
301
138, 172, 200, 201, 218, 223, 229, 235, 237, 262,
shear rate, 128, 129, 130, 131, 134, 141, 262, 263,
264, 265, 266, 280, 282
275
testing, 66, 145, 178, 242, 267
sludge, vii, ix, xiv, 5, 13, 14, 18, 21, 22, 29, 34, 37,
toxic effect, 198
47, 48, 49, 50, 63, 70, 99, 100, 101, 102, 103,
toxic metals, 49
104, 133, 144, 177, 218, 220, 222, 223, 239, 242,
toxic substances, x, 143
243, 250, 253, 257, 265, 266, 267, 269, 274
toxicity, 5, 20, 27, 28, 70, 138, 220, 222, 235, 266,
sodium, 19, 27, 40, 44, 50, 72, 86, 122, 134, 144,
269, 271
158, 190, 195, 201, 211, 225, 263, 268, 270, 271,
toxin, 199, 203, 208, 209
273, 282, 298
transition metal, 278, 292
sodium hydroxide, 72, 158, 190, 201, 298
transition metal ions, 278, 292
solid phase, 157, 291
turbidity, vi, viii, ix, xi, xii, xiii, xiv, 16, 17, 28, 29,
solid state, 22, 292
30, 31, 35, 38, 42, 47, 58, 59, 60, 61, 62, 63, 64,
solid waste, xiii, 144, 217, 218, 223
69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 82, 91,
solubility, xv, 11, 24, 43, 118, 221, 223, 262, 263,
120, 126, 128, 153, 155, 156, 157, 160, 163, 164,
264, 267, 268, 270, 277, 280, 284, 285, 290
165, 166, 170, 171, 172, 173, 175, 179, 181, 183,
solution, xi, xii, xv, 4, 5, 7, 16, 25, 28, 35, 48, 58, 72,
184, 185, 194, 197, 199, 200, 201, 202, 203, 204,
74, 78, 83, 85, 87, 88, 120, 125, 149, 155, 157,
205, 206, 207, 208, 210, 212, 213, 218, 220, 225,
158, 159, 165, 181, 184, 187, 194, 200, 208, 219,
230, 231, 234, 238, 239, 243, 244, 245, 246, 253,
225, 227, 250, 257, 263, 266, 273, 275, 277, 278,
259, 260, 262, 266, 267, 268, 270, 272, 278, 279,
279, 280, 281, 282, 285, 286, 289, 290, 296, 298
280, 284, 285, 295, 296, 297, 298, 299, 300, 301,
solvents, 11, 12, 24, 291
302
species, 21, 28, 49, 50, 51, 52, 53, 54, 59, 64, 89,
109, 110, 111, 112, 113, 114, 115, 120, 121, 124,
131, 135, 136, 212, 221, 251, 263, 264, 265, 267, V
269
specific adsorption, 83, 84 velocity, 87, 89, 90, 91, 94, 95, 96, 97, 128, 133,
spectroscopy, xv, 273, 277 140, 157, 262, 269, 296, 297
stability, ix, 8, 15, 17, 20, 23, 47, 49, 82, 122, 174, viscosity, 14, 96, 264, 265, 266
261, 273, 288, 292 volatile organic compounds, 229
stabilization, 9, 118, 121, 222, 268
starch, 2, 3, 12, 13, 16, 17, 22, 27, 28, 29, 30, 32, 33, W
34, 35, 37, 39, 40, 42, 43, 111, 116, 124, 125,
129, 266, 267, 269, 273, 290 waste management, 48, 218, 219
starch granules, 39 wastewater treatment, viii, ix, xi, xiv, xv, 2, 6, 13,
stock, 58, 200, 280, 281, 282, 298 17, 18, 21, 24, 30, 32, 33, 35, 36, 39, 40, 42, 43,
310 Index
44, 65, 66, 79, 81, 82, 89, 97, 99, 101, 102, 103, water-soluble polymers, 104
104, 118, 125, 133, 169, 170, 171, 172, 174, 178, waterways, 103
180, 227, 259, 262, 268, 271, 275, 291, 292
water purification, 138, 188
X
water quality, viii, x, xiii, 71, 82, 105, 107, 108, 110,
138, 142, 170, 172, 197, 198, 262, 267
xanthan gum, 27, 29
water shortages, 219
water treatment, viii, x, xii, xiii, 23, 26, 32, 43, 44,
48, 49, 63, 64, 65, 66, 69, 70, 73, 78, 79, 82, 91, Z
102, 104, 107, 108, 110, 111, 115, 120, 123, 126,
133, 134, 137, 138, 140, 141, 142, 144, 156, 162, zeta potential, xi, 6, 61, 62, 115, 169, 170, 171, 172,
165, 166, 167, 187, 188, 193, 195, 198, 208, 212, 173, 174, 175, 176, 178, 220, 222, 279, 280, 282,
213, 214, 215, 216, 222, 251, 254, 255, 258, 266, 285, 286, 289, 292
267, 272, 273, 274, 275, 278, 295, 301, 302