Вы находитесь на странице: 1из 49

ARTICLE IN PRESS

Prog. Polym. Sci. 33 (2008) 399447


www.elsevier.com/locate/ppolysci

Application of chitosan, a natural aminopolysaccharide, for dye


removal from aqueous solutions by adsorption processes using
batch studies: A review of recent literature
Gregorio Crini, Pierre-Marie Badot
Department of Chrono-Environment, University of Franche-Comte, UMR UFC/CNRS 6565, Place Leclerc, 25000 Besanc- on, France
Received 21 December 2006; received in revised form 9 November 2007; accepted 9 November 2007
Available online 17 November 2007

Abstract
Application of chitinous products in wastewater treatment has received considerable attention in recent years in the
literature. In particular, the development of chitosan-based materials as useful adsorbent polymeric matrices is an
expanding eld in the area of adsorption science. This review highlights some of the notable examples in the use of chitosan
and its grafted and crosslinked derivatives for dye removal from aqueous solutions. It summarizes the key advances and
results that have been obtained in their decolorizing application as biosorbents. The review provides a summary of recent
information obtained using batch studies and deals with the various adsorption mechanisms involved. The effects of
parameters such as the chitosan characteristics, the process variables, the chemistry of the dye and the solution conditions
used in batch studies on the biosorption capacity and kinetics are presented and discussed. The review also summarizes and
attempts to compare the equilibrium and kinetic models, and the thermodynamic studies reported for biosorption onto
chitosan.
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Chitosan; Biosorption; Dyes; Batch process; Modeling and thermochemistry of biosorption

Abbreviation: AB, acid blue; AB 1, acid black 1; AB 15, acid blue 15; AB 25, acid blue 25; AB 40, acid blue 40; AB 62, acid blue 62; AB
113, acid blue 113; AG 25, acid green 25; AG 27, acid green 27; AO 7, acid orange 7; AO 10, acid orange 10; AO 12, acid orange 12; AO
51, acid orange 51; AR, acid red; AR 1, acid red 1; AR 14, acid red 14; AR 18, acid red 18; AR 73, acid red 73; AR 27, acid red 27; AR 87,
acid red 87; AR 88, acid red 88; AR 138, acid red 138; AV 5, acid violet 5; AY 25, acid yellow 25; BB, basic blue; BB 1, basic brown 1; BB
3, basic blue 3; BB 9, basic blue 9; BR, brilliant red M5BR2; BY 45, basic yellow 45; CV, crystal violet; DB, direct blue; DB 14, direct blue
14; DB 71, direct blue 71; DO, direct orange; DR, direct red; DR 2, direct red 2; DR 81, direct red 81; DS, direct scarlet B; DY 4, direct
yellow 4; IC, indigo carmine; IR, iragalon rubine RL; MB, maxilon blue 4GL; MB 29, mordant blue 29; MB 33, mordant brown 33; MO,
methyl orange; MO 10, mordant orange 10; MY, metanil yellow; MY 30, mordant yellow 30; O II, orange II; Rb 5, reactive blue 5; RB,
reactive blue RN; RB 5, reactive black 5; RB 2, reactive blue 2; RB 15, reactive blue 15; RB 19, reactive blue 19; RB 222, reactive blue 222;
RO, reactive orange; RO 16, reactive orange 16; R 6G, rhodamine 6G; RR, reactive red; RR B, reactive red RB; RR 2, reactive red 2; RR
141, reactive red 141; RR 189, reactive red 189; RR 195, reactive red 195; RR 222, reactive red 222; RTB, reactive T-blue; RY, reactive
yellow GR; RY 2, reactive yellow 2; RY 86, reactive yellow 86; RY 145, reactive yellow 145.
Corresponding author. Tel.: +33 3 81 66 57 01; fax: +33 3 81 66 57 97.
E-mail address: gregorio.crini@univ-fcomte.fr (G. Crini).
0079-6700/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2007.11.001

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

400

Contents
1.
2.

3.
4.

5.
6.

7.
8.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400
General considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
2.1. Batch experiments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
2.2. Why to use chitosan as raw material? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
2.3. Considerations on dye adsorption. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
2.4. Why to use chitosan as a biosorbent for dye removal? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
2.5. Raw chitosan and chitosan-based materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
A brief review of the recent literature on the adsorption of dyes by chitosan . . . . . . . . . . . . . . . . . . . . . . . 412
Control of adsorption performances of chitosan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
4.1. Inuence of the chitosan characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
4.1.1. Chitosan origin. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
4.1.2. Physical nature of the chitosan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
4.1.3. Chemical structure of chitosan. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
4.2. Activation conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
4.2.1. Chitosan preprotonation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
4.2.2. Grafting reactions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
4.2.3. Inuence of crosslinking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
4.2.4. Chitosan-based composite beads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
4.3. Inuence of process variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
4.3.1. Effect of chitosan dosage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
4.3.2. Effect of initial dye concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
4.3.3. Effect of contact time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
4.3.4. Effect of stirring rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
4.3.5. Effect of dryness. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
4.4. Chemistry of the dye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424
4.5. Effect of the solution conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
4.5.1. Effect of pH. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 425
4.5.2. Effect of pH variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
4.5.3. pH sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
4.5.4. Effect of ionic strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
4.5.5. Effect of competitive molecules and ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
4.6. Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
4.7. Desorption of dyes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
Adsorption mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 429
Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
6.1. Equilibrium isotherm models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
6.2. Kinetic modeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
6.3. Thermochemistry of biosorption. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
6.3.1. Effect of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
6.3.2. Thermodynamic parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
Economic aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
Concluding remarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 444

1. Introduction
Many industries, such as textile, paper, plastics
and dyestuffs, consume substantial volume of water,
and also use chemicals during manufacturing and
dyes to color their products. As a result, they
generate a considerable amount of polluted waste-

water [15]. For example, pulp and paper mills


generate varieties of pollutants depending upon the
type of the pulping process. Their toxic efuents
are a major source of aquatic pollution and will
cause considerable damage to the receiving waters if
discharged untreated [1]. This specic type of
pollution is characterized by high biochemical

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

Nomenclature

k2

aL
C

ki

Ce
Co
DG
DH
DS
Ea
KF
KL
k0
k1

Langmuir isotherm constant (l/mg)


intercept of the intraparticle diffusion
equation (mg/g)
liquid-phase dye concentration at equilibrium (mg/l)
initial dye concentration in liquid phase
(mg/l)
Gibbs free energy change (kJ/mol)
enthalpy change (kJ/mol)
entropy change (J/mol K)
activation energy (kJ/mol)
Freundlich isotherm constant (l/g)
Langmuir isotherm constant (l/g)
frequency factor (min1)
equilibrium rate constant of pseudo-rstorder adsorption (min1)

oxygen demand (BOD), chemical oxygen demand


(COD), suspended solids (mainly bers), bad smell,
toxicity (high concentration of nutrients, presence
of chlorinated phenolic compounds, sulfur and
lignin derivatives, etc.), and especially color [1,2].
Color is the rst contaminant to be recognized
in wastewater and the presence of very small
amounts of dyes in water is highly visible and
undesirable [4,5].
During the past three decades, several wastewater
treatment methods have been reported and attempted for the removal of pollutants from textile,
pulp and paper mill efuents. The technologies can
be divided into three main categories: (i) conventional methods, (ii) established recovery processes
and (iii) emerging removal methods (see Table 1). In
the literature, there are a great number of feasibility
studies concerning the treatment of dyeing efuents
by these methods [28].
It is known that wastewaters containing dyes are
very difcult to treat, since the dyes are recalcitrant
molecules (particularly azo dyes), resistant to
aerobic digestion, and are stable to oxidizing agents.
Another difculty is treatment of wastewaters
containing low concentrations of dye molecules. In
this case, common methods for removing dyes are
either economically unfavorable and/or technically
complicated. Because of the high costs associated
with their practical applications to remove trace
amounts of impurities, many of the methods for
treating dyes in wastewater (Table 1) have not been

qe
qt
qmax
m
nF
R
T
t
te
V
x

401

equilibrium rate constant of pseudosecond-order adsorption (g/mg min)


intraparticle diffusion rate constant
(mg/g min1/2)
amount of dye adsorbed at equilibrium
(mg/g)
amount of dye adsorbed at time t (mg/g)
maximum adsorption capacity of the
adsorbent (mg/g)
mass of adsorbent used (g)
Freundlich isotherm exponent
universal gas constant (8.314 J/mol K)
absolute temperature (1K)
time (min)
equilibrium time (min)
volume of dye solution (l)
amount of dye adsorbed (mg)

widely applied on a large scale in the paper and


textile industries. In practice, no single process
provides adequate treatment and a combination of
different processes is often used to achieve the
desired water quality in the most economical way.
Thus, there is a need to develop new decolorization
methods that are effective and acceptable in
industrial use.
It is now recognized that adsorption using
low-cost adsorbents is an effective and economic
method for water decontamination. A large variety
of non-conventional adsorbents materials have been
Table 1
Principal existing and emerging processes for dyes removal
Conventional treatment
processes

 Coagulation/oculation
 Precipitation/oculation
 Electrocoagulation/
electrootation

 Biodegradation
 Adsorption on activated carbon
Established removal
methods







Emerging recovery
technologies

 Advanced oxidation
 Selective bioadsorption
 Biomass

Oxidation
Electrochemical treatment
Membrane separation
Ion-exchange
Incineration

ARTICLE IN PRESS
402

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

proposed and studied for their ability to remove


dyes [6]. However, low-cost adsorbents with high
adsorption capacities are still under development to
reduce the adsorbent dose and minimize disposal
problems. Much attention has recently been focused
on various biosorbent materials such as fungal or
bacterial biomass and biopolymers that can be
obtained in large quantities and that are harmless to
nature. Special attention has been given to polysaccharides such as chitosan, a natural aminopolymer. It is clear from the literature that the
biosorption of dyes using chitosan is one of the
more frequently reported emerging methods for
the removal of pollutants.
Chitosan has been investigated by several researchers as a biosorbent for the capture of
dissolved dyes from aqueous solutions. This natural
polymer possesses several intrinsic characteristics
that make it an effective biosorbent for the removal
of color. Its use as a biosorbent is justied by two
important advantages: rstly, its low cost compared
to commercial activated carbon (chitosan is derived
by deacetylation of the naturally occurring biopolymer chitin which is the second most abundant
polysaccharide in the world after cellulose); secondly, its outstanding chelation behavior (one of the
major applications of this aminopolymer is based
on its ability to tightly bind pollutants, in particular
heavy metal ions).
In this paper, we review the use of chitosan for
dye removal from aqueous solutions. Since the
review only presents data obtained using raw,
grafted and crosslinked chitosans, the discussion
will be limited to these chitosan-based materials and
their adsorption properties. The main objectives are
to summarize some of the developments related to
the decolorizing applications of these polymeric
materials and to provide useful information about
their most important features. We give an overview
of several recent batch studies reported in the
literature, with the various mechanisms involved.
To do so, an extensive list of recent literature has
been compiled. The effects of various parameters
such as chitosans characteristics, the activation
conditions, the process variables, the chemistry of
the dye and the experimental conditions used in
batch systems, on biosorption are presented and
discussed. The review also summarizes the equilibrium and kinetic models, and the thermodynamic
studies reported for biosorption onto chitosan,
which are important to determine the biosorption
capacity and to design treatment processes.

2. General considerations
2.1. Batch experiments
The change in the concentration of a pollutant
(adsorbate) in the surface layer of the material
(adsorbent) in comparison with the bulk phase with
respect to unit surface area is termed adsorption.
The term biosorption is given to adsorption
processes, which use biomaterials as adsorbents
(or biosorbents). The assessment of a solid-liquid
adsorption system is usually based on two types of
investigations: batch adsorption tests and dynamic
continuous-ow adsorption studies. The present
review only presents data obtained using batch
studies. When studying adsorption from solutions
on materials it is convenient to differentiate between
adsorption from dilute solution and adsorption
from binary and multicomponent mixtures covering
the entire mole fraction scale. To judge by the
number of papers published annually on adsorption
from dilute solution, this subject is more important
than adsorption from binary mixtures. Therefore,
reference will be made hereafter to adsorption from
dilute aqueous solutions.
Batch studies use the fact that the adsorption
phenomenon at the solid/liquid interface leads to a
change in the concentration of the solution.
Adsorption isotherms are constructed by measuring
the concentration of adsorbate in the medium
before and after adsorption, at a xed temperature.
For this, in general, adsorption data including
equilibrium and kinetic studies are performed using
standard procedures consisting of mixing a xed
volume of dye solution with an known amount of
chitosan in controlled conditions of contact time,
agitation rate, temperature and pH. At predetermined times, the residual concentration of the dye is
determined by spectrophotometry at the maximum
absorption wavelength. Dye concentrations in solution can be estimated quantitatively using linear
regression equations obtained by plotting a calibration curve for each dye over a range of concentrations. The adsorption capacity (adsorption uptake
rate) is then calculated and is usually expressed in
milligrams of dye adsorbed per gram of the (dry)
adsorbent. For example, the amount of dye
adsorbed at equilibrium, qe, is calculated from
the mass balance equation given by Eq. (1). The
symbols used in the equation are dened in the
Nomenclature section. In general, the experiments are conducted in triplicate under identical

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

403

Simplicity, well-established experimental methods,


and easily interpretable results are some of the
important reasons frequently evoked for the extensive usage of these methods. Another interesting
advantage is the fact that, in batch systems, the
parameters of the solution such as adsorbent
concentration, pH, ionic strength, temperature,
etc. can be controlled and/or adjusted.

conditions and found reproducible:


V C o  C e
.
(1)
m
The equilibrium relationship between adsorbent
and adsorbate, i.e. the distribution of dye molecules
between the solid adsorbent phase and the liquid
phase at equilibrium, which are the basic requirements for the design of adsorption systems, are
described by adsorption isotherms using any of the
mathematical models available. The equilibrium
adsorption isotherm, usually the ratio between the
quantity adsorbed and that remaining in solution at
a xed temperature at equilibrium, is fundamentally
important since the equilibrium studies give the
capacity of the adsorbent and describe the adsorption isotherm by constants whose values express the
surface properties and afnity of the adsorbent (i.e.
to study the interaction between the adsorbate and
the surface and to know about the structure of the
adsorbed layer).
In the literature, batch methods are widely used
to describe the adsorption capacity and the adsorption kinetics. These processes are cheap and simple
to operate and, consequently, often favoured for
small- and medium-size process applications using
simple and readily available mixing tank equipment.

qe

2.2. Why to use chitosan as raw material?


The majority of commercial polymers and ionexchange resins are derived from petroleum-based
raw materials using processing chemistry that is not
always safe or environmental friendly. Today, there
is growing interest in developing natural low-cost
alternatives to synthetic polymers [6].
Chitin, found in the exoskeleton of crustaceans,
the cuticles of insects, and the cells walls of fungi, is
the most abundant aminopolysaccharide in nature
[911]. This low-cost material is a linear homopolymer composed of b(1-4)-linked N-acetyl glucosamine (Fig. 1). It is structurally similar to cellulose,
but it is an aminopolymer and has acetamide groups
at the C-2 positions in place of the hydroxyl groups.
The presence of these groups is highly advantageous,

CH2OH

CH2OH
O

O
O

OH

OH
n

n
NHCOCH3

NH2

Chitin

Chitosan

CH2OH

NH2
O
O

OH

OH

O
O

DA
NHCOCH3

1-DA

CH2OH

N-acetyl glucosamine unit

glucosamine unit
Commercial Chitosan

Fig. 1. Chemical structure of chitin [poly(N-acetyl-b-D-glucosamine)], chitosan [poly(D-glucosamine)] and commercial chitosan (a copolymer
characterized by its average degree of acetylation (DA)).

ARTICLE IN PRESS
404

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

Partial deacetylation of chitin results in the


production of chitosan (Fig. 2), which is a
polysaccharide composed by polymers of glucosamine and N-acetyl glucosamine. The chitosan
label generally corresponds to polymers with less
than 25% acetyl content. The fully deacetylated
product is rarely obtained due to the risks of side
reactions and chain depolymerization. Copolymers
with various extents of deacetylation and grades are
now commercially available. Chitosan and chitin
are of commercial interest due to their high
percentage of nitrogen compared to synthetically
substituted cellulose. Chitosan is soluble in acid
solutions and is chemically more versatile than
chitin or cellulose. The main reasons for this are
undoubtedly its appealing intrinsic properties, as
documented in a recent review [11], such as
biodegradability, biocompatibility, lm-forming
ability, bioadhesivity, polyfunctionality, hydrophilicity and adsorption properties (Table 2). Most of
the properties of chitosan can be related to its
cationic nature [912], which is unique among
abundant polysaccharides and natural polymers.
These numerous properties lead to the recognition
of this polyamine as a promising raw material for
adsorption purposes.

providing distinctive adsorption functions and


conducting modication reactions. The raw polymer is only commercially extracted from marine
crustaceans primarily because a large amount of
waste is available as a by-product of food processing [9]. Chitin is extracted from crustaceans
(shrimps, crabs, squids) by acid treatment to
dissolve the calcium carbonate followed by alkaline
extraction to dissolve the proteins and by a
decolorization step to obtain a colorless product
[10,11] (Fig. 2).
Since the biodegradation of chitin is very slow in
crustacean shell waste, accumulation of large
quantities of discards from processing of crustaceans has become a major concern in the seafood
processing industry. So, there is a need to recycle
these by-products. Their use for the treatment of
wastewater from another industries could be helpful
not only to the environment in solving the solid
waste disposal problem, but also to the economy.
However, chitin is an extremely insoluble material.
Its insolubility is a major problem that confronts the
development of processes and uses of chitin [11],
and so far, very few large-scale industrial uses have
been found. More important than chitin is its
derivative, chitosan (Fig. 1).

Shellfish wastes
demineralization
deproteinization
decoloration
hydrolysis

carb oxymethylation
Chitin

glucosamines
oligosaccharides

deacetylation

carb oxymethylchitin

Chitosan
derivatization

oligosaccharides
glucosamines

salts

chitosan derivatives

acetylation
N-acetyl-D-glucosamines
Fig. 2. Simplied representation of preparation of chitin, chitosan and their derivatives.

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
Table 2
Intrinsic properties of chitosan
Physical and
chemical properties

 Linear aminopolysaccharide with


high nitrogen content

 Rigid D-glucosamine structure; high


crystallinity; hydrophilicity

 Capacity to form hydrogen bonds


intermolecularly; high viscosity

 Weak base; the deprotonated amino







Polyelectrolytes (at
acidic pH)

 Cationic biopolymer with high







Biological
properties

group acts a powerful nucleophile


(pKa 6.3)
Insoluble in water and organic
solvents; soluble in dilute aqueous
acidic solutions
Numerous reactive groups for
chemical activation and crosslinking
Forms salts with organic and
inorganic acids
Chelating and complexing properties
Ionic conductivity

charge density (one positive charge


per glucosamine residue)
Flocculating agent; interacts with
negatively charged molecules
Entrapment and adsorption
properties; ltration and separation
Film-forming ability; adhesivity
Materials for isolation of
biomolecules

 Biocompatibility
J
J
J

Non-toxic
Biodegradable
Adsorbable

 Bioactivity
J
J
J
J

Antimicrobial activity (fungi,


bacteria, viruses)
Antiacid, antiulcer, and
antitumoral properties
Blood anticoagulants
Hypolipidemic activity

 Bioadhesivity

The interest in chitin and chitosan is reected by


an increasing number of articles published (Fig. 3),
and of meetings in Europe, Asia and America on
this topic. Table 3 summarizes the main applications of chitin and chitosan. Currently, these
polymers and their numerous derivatives are widely
used in pharmacy [21,36,37], medicine [11,21,2329],
biotechnology [10,21,30], chemistry [21,3134], cosmetics and toiletries [11,21], food technology [35],
and the textile [21], agricultural [12,20,21], pulp and

405

paper industries [21] and other elds [21,38,39] such


as enology, dentistry and photography. The potential industrial use of chitosan is widely recognized.
These versatile materials are also widely applied in
clarication and water purication, and water and
wastewater treatment as coagulating [1315], occulating [16,17] and chelating agents [1922]. However, despite a large number of studies on the use of
chitosan for pollutant recovery in the literature, this
research eld has failed to nd practical applications on the industrial scale: this aspect will be
discussed later.
2.3. Considerations on dye adsorption
Synthetic dyes are an important class of recalcitrant organic compounds and are often found in the
environment as a result of their wide industrial use.
These industrial pollutants are common contaminants in wastewater and are difcult to decolorize
due to their complex aromatic structure and
synthetic origin. They are produced on a large
scale. Although the exact number (and also the
amount) of the dyes produced in the world is not
known, there are estimated to be more than 100,000
commercially available dyes. Many of them are
known to be toxic or carcinogenic.
Generally, dyes can be classied with regard to
their chemical structure (e.g. azo, anthraquinone,
indigo, triphenylmethane), with regard to the
method and domain of usage (e.g. direct, reactive,
chromic, metal-complexes, disperse, mordant, sulfur, vat, pigments), and/or with regard to their
chromogen (e.g. n-p*, donoracceptor, cyanine,
polyenes). Mishra and Tripathy [40] proposed a
simplied classication as follows: anionic (direct,
acid and reactive dyes), cationic (basic) dyes and
non-ionic (disperse) dyes. As mentioned, there are
many structural varieties such as acidic, disperse,
basic, azo, diazo, anthraquinone-based and metal
complex dyes. Azo and anthraquinone colorants are
the two major classes of synthetic dyes and
pigments. Together they represent about 90% of
all organic colorants.
Fig. 4 gives some examples of dyes currently used
in the textile industry. Reactive Black 5, a diazo dye,
has two sulfonate groups and two sulfatoethylsulfon groups in its molecular structure that have
negative charges in aqueous solution. Basic Blue 3, a
monoxazine dye, possesses an overall positive
charge because it tends to ionize in solution. The
anthraquinonic dyes Reactive Blue 19 and Disperse

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

406

300

Number of articles

250
200
150
100
50
0
1998

1999

2000

2001

2002

2003

2004

2005

4%
7%

coagulation

3%

precipitation
flocculation
adsorption

53%

28%
flotation
filtration
membranes

1%
4%

Fig. 3. A Scopus database literature survey of the wastewater applications of chitosan and chitin: (a) research articles published from 1998
to 2005 (the survey did not include patents) and (b) main domains of chitosan and chitin in the removal of pollutants from solutions.

Blue 14 have an anionic and non-ionic character,


respectively. Basic Green 4 is an N-methylated
diaminotriphenyl methane dye, which has a cationic
character. It is important to note that dye molecules
have many different and complicated structures,
and their adsorption behavior is directly related to
the chemical structure, the dimensions of the dye
organic chains, and the number and positioning of
the functional groups of the dyes. This is one of the
most important factors inuencing adsorption.
However, to the weay adsorption is affected by
the chemical structure of the dyes was not clearly
identied: this aspect will be discussed in the
following sections.
Generally, a suitable adsorbent for adsorption
process of dye molecules should meet several
conditions:




low cost,
readily available,





large capacity and rate of adsorption,


high selectivity for different concentrations,
and efcient for removal of a wide variety of
target dyes.

Recently, numerous low-cost adsorbents have


been proposed for dye removal. Among them,
non-conventional activated carbons from solid
wastes, industrial by-products, agricultural solid
wastes, clays, zeolites, peat, polysaccharides and
fungal or bacterial biomass deserve particular
attention as recently summarized in a review by
Crini [6]. Each has advantages and drawbacks.
However, at the present time, there is no single
adsorbent capable of satisfying the above requirements. Thus, there is a need for new systems to be
developed. In addition, the adsorption process
provides an attractive alternative treatment, especially if the adsorbent is selective and effective for
removal of anionic, cationic and non-ionic dyes.

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
Table 3
Applications of chitin and chitosan
Fields

Applications

Agriculture

Protection of plants
Increase of crop yields (reduces the growth
of phytopathogenic fungi)
Seed and fertilizer coating; soil treatment

Biomedical
engineering

Biological activities (antifungal,


antimicrobial, antiinfectious); antitumor
agent
Hemostatic effects; enhances blood
coagulation
Promotes tissue growth; stimulates cell
proliferation; articial skin
Sutures/bandages
Ophthalmology, contact lenses

Biotechnology

Enzyme and cell immobilization


Cell-stimulating materials
Matrix for afnity chromatography or
membranes

Chemical
industry

Water purication (metal chelation); water


engineering (occulation, ltration,
adsorption); sludge treatment
Reverse osmosis, ltration membranes; gas
separation
Production of biodegradable packaging
lms
Catalysis

Cosmetics and
toiletries

Hair spray, lotion; hand and body creams;


shampoo, moisturizer

Food industry

Diet foods and dietary ber;


hypocholesterolemic activity (binds
cholesterol, fatty acids and
monoglycerides)
Preservation of foods from microbial
deterioration
Bioconversion for the production of valueadded food products
Recovery of waste material from foodprocessing discards
Clarication and deacidication of fruit
juices and beverages
Emulsifying agent; colour stabilization
Animal feed additive

Pharmaceutics

Controlled drug delivery carriers


Microcapsules (forming gels and capsules
with anionic polymers)
Dermatological products (treats acne)

Others

Textiles (anti-bacterial properties)


Pulp and paper (wet strength)
Enology (clarication, deacidication)
Dentistry (dental implants)
Photography (paper)

407

Now, the amounts of dyes adsorbed on the above


adsorbents are not very high, some have capacities
between 100 and 600 mg/g and some even lower
than 50 mg/g [6]. To improve the efciency and
selectivity of the adsorption processes, it is essential
to develop more effective and cheaper adsorbents
with higher adsorption capacities.
2.4. Why to use chitosan as a biosorbent for dye
removal?
As already mentioned, a growing number of
papers have been published since the 1980s concerning chitosan for wastewater treatment. In
particular, chitosan has received considerable interest in heavy metal chelation due to its relatively low
cost compared with commercial activated carbon,
its excellent metal-binding capacities and interesting
selectivity, as well as its possible biodegradability
after use. It is frequent to reach adsorption
capacities as high as 3 mmol metal per gram
chitosan for Cu (i.e. 200 mg/g), 12 mmol metal
per gram for Pt and Pd, and up to 710 mmol metal
per gram for Mo and V [18,19]. In accordance with
the very abundant data in the literature, liquidphase adsorption using chitosan is one of the most
popular methods for the removal of heavy metals
from wastewater since proper design of the adsorption process will produce a high-quality treated
solution. Readers interested in a detailed discussion
of the interaction of metal ions with chitosan should
refer to the excellent comprehensive review by
Guibal [18].
Besides being natural and plentiful, chitosan
possesses interesting characteristics that also make
it an effective biosorbent for the removal of color
with outstanding adsorption capacities. Compared
with conventional commercial adsorbents such as
commercial activated carbons (CAC) for removing
dyes from solution, adsorption using chitosan-based
materials as biosorbents offers several advantages
(Table 4). In particular, three factors have specically contributed to the growing recognition of
chitosan as a suitable biomaterial for dye removal:

First is the fact that the chitosan-based polymers


are low-cost materials obtained from natural
resources and their use as biosorbents is extremely cost-effective. In many countries, shery
wastes were used as excellent sources to produce
chitosan. Since such waste is abundantly available, chitosan may be produced at relatively low

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

408

NHCH3

NHCH3

Disperse Blue 14
N

NaO3SOCH2CH2O2S

SO3Na

HO
Reactive Black 5
H2N
N
NaO3SOCH2CH2O2S

SO3Na
(C2H5)2N

Cl+
N(C2H5)2

O
Basic Blue 3

NH2
N+

SO3Na

HO

-O

Basic Green 4
O

HN

Reactive Blue 19

SO2CH2CH2OSO3Na
N(CH3)2

Fig. 4. Examples of commonly used dyestuffs in the textile industry.

cost. The volume of biosorbent used is also


reduced as compared to conventional adsorbents
since they are more efcient.
Second is the high adsorption capacities reported. The biosorbents posses an outstanding
capacity and high rate of adsorption, and also
high selectivity in detoxifying both very diluted
or concentrated solutions. They also have an
extremely high afnity for many varieties of dyes.
The third factor is the development of new
complexing materials because chitosan is versatile: it can be manufactured into lms, membranes, bers, sponges, gels, beads and
nanoparticles, or supported on inert materials.
The utilization of these materials presents many
advantages in terms of applicability to a wide
variety of process congurations.

Of course, there are, also disadvantages of using


chitosan in wastewater treatment (Table 4). This
research eld fails to nd practical application at the
industrial scale. There are several reasons for
explaining this difculty in transferring the process

to industrial applications [10,11,18,20]. The adsorption properties depend on the different sources of
chitin (the quality of commercial chitin available is
not uniform) and performance is also dependent on
the type of material used. Another important
criterion to be taken into account concerns the
variability and heterogeneity of the polymer (the
difculty of controlling the distribution of the acetyl
groups along the backbone makes it difcult to get
reproducible initial polymers). There is a need for a
better standardization of the production process to
be able to prepare reproducible initial polymers
having the same characteristics. Changes in the
specications of the polymer may signicantly
change adsorption performance. Another problem
with chitosan derivatives is their poor physicochemical characteristics, in particular low surface area
and porosity. In addition, although chitosan is
much easier to process than chitin or other low-cost
adsorbents, the stability of chitosan materials is
generally lower, owing to their more hydrophilic
character and, especially, pH sensitivity. Being a
biopolymer, chitosan is biodegradable and this may

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
Table 4
Advantages and disadvantages of chitosan and chitosan-based
materials used as biosorbent for the removal of dyes from
aqueous solutions
Advantages

 Low-cost hydrophilic









biopolymer
Very abundant material
and widely available in
many countries
Renewable resource
Cationic polysaccharide (in
acidic medium)
Environmentally friendly,
publicly acceptable
material
Extremely cost effective
Outstanding dye-binding
capacities of a wide range
of dyes
Fast kinetics
High selectivity in
decolorizing both very
dilute or concentrated
solutions
Versatile biosorbent

Disadvantages

 Variability in the polymer


characteristics

 The performance depends







of the origin and treatment


of the polymer, and also its
degree of N-acetylation
Nonporous sorbent
Requires chemical
derivatization to improve
its performance
Not effective for cationic
dyes (except after
modication)
pH sensitivity
Its use in sorption columns
is limited (hydrodynamic
limitations and column
fouling)
Non-destructive process

be a serious drawback for long-term applications.


These problems can rebut industrial users. Readers
interested in a detailed discussion of these problems
should refer to the work of Guibal [18]. However,
the opportunity now exists to consider chitosan for
emerging applications where other technologies
would be unsuitable.
Different reviews of chitosan-based biomaterials
have been reported concerning adsorption and
separation, including metal complexation [18,19],
complexing adsorbent matrices [21,22,41,42], and
membranes [33]. Obviously, chitosan has also been
investigated as a biosorbent for the capture of
dissolved dyes from aqueous solutions in numerous
articles. The effectiveness of chitin and chitosan to
adsorb dye molecules has been reported by numerous workers [4357]. For example, as long ago as
1958, Giles et al. [43] investigated the binding
behavior of dyes to chitin. In 19821985, extensive
studies on the adsorption of dyes on chitin by
McKay et al. [4448] also revealed that chitin can
adsorb substantial quantities of dyestuffs from
aqueous solutions. The interaction of chitosan with
dyes was studied by several workers [4957]. These
earlier papers clearly demonstrated that raw materials have an intrinsically high afnity and selectivity

409

for a wide range of dyes, although several contradictory observations have been reported. However,
a few review articles on the potential of chitosan for
dye removal have been published. The application
of the adsorption of pollutants including dyes onto
chitosan has been reviewed by Ravi Kumar [21] and
No and Meyers [22]. Various chitosan-based composites and membranes have been also developed
and proposed for adsorption and separation purposes [33,42]. To avoid repetition, in the following
chapters, only raw, grafted and crosslinked chitosans will be discussed. This review focuses on the
recent developments related to decolorizing applications of the chitosan-based materials and reports the
main advances published over the last 10 years. This
is an ambitious project since the very large number
of groups working around the world forces us to
make a selection from the most signicant results.
Table 5 lists some of the researchers whose results
are discussed in this review and the dyes they
investigated [58116].
2.5. Raw chitosan and chitosan-based materials
Practical use of chitosan has been mainly
conned to the unmodied forms. For a breakthrough in its utilization, chemical derivatization
onto polymer chains has been proposed to produce
new materials. Derivatization is a key point which
will introduce the desired properties to enlarge the
eld of its potential applications. Chitosan has three
types of reactive functional groups, an amino group
as well as both primary and secondary hydroxyl
groups at the C-2, C-3 and C-6 positions, respectively (Fig. 1). Its advantage over other polysaccharides is that its chemical structure allows specic
modications without too many difculties, especially, at the C-2 position [11]. These functional
groups allow direct substitution reactions and
chemical modications, yielding numerous useful
materials for different domains of application.
The most commonly used chemical activations are
carboxymethylation, acetylation and grafting. The
variety of groups which can be attached to
the polymer is almost unlimited. To control both
the physical, mechanical and chemical properties,
various techniques can be used, and often, the
methods are adapted from the cellulose world [11].
The chitosan derivatives can be classied into four
main classes of materials: modied polymers, crosslinked chitosans, chitosan-based composites and
membranes (Table 6).

ARTICLE IN PRESS
410

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

Table 5
Authors of recent research on dye removal by chitosan (selected papers)
Corresponding author

Country

Dye(s)

Reference(s)

Airoldi C.
Annadurai G.
Cestari AR.
Chen DH.
Chen L.
Chiou MS.

Brazil
Iran
Brazil
Taiwan
China
Taiwan

[58]
[59,60]
[6163]
[64]
[65]
[6670]

Cho SY.
Crini G.
de Favere VT.
Dutta PK.
El-Tahlawy KF.
Fahmy HM.
Guibal E.

Korea
France
Brazil
India
Egypt
Egypt
France

Guha AK.
Hebeish R
Juang RS.
Li HY.
Martel B.
Manolova N.
McKay G.
Miyata K.
Prado AGS.
Saha TK.
Shimizu Y.
Shyu SS.
Stevens WF.
Thiravetyan P.
Szeto YS.
Uzun I.
Wen YZ.

India
Egypt
Taiwan
Taiwan
France
Bulgaria
Hong Kong
Japan
Brazil
Bangladesh
Japan
Taiwan
Thailand
Thailand
Hong Kong
Turkey
China

BB 9
BB 9, DS
IC, RB, RN, RR, RY
AG 25, AO 12
AB, BB
AO 7, AO 12, AR 14, DR 81 MY, RB 2, RB 15, RR 2,
RR 189, RR 222, RY 2, RY 86
RB 5
BB 3, BB 9
RO 16
DB
BR, IR, MB
DR
AB 1, AB 113, AG 25, AV 5, AY 25, DB 14, DB 71,
DY 4, MB 29, MB 33, MO 10, RB 5
AR 87
AR, BY 45, DO, RO
AO 51, BB 9, RB 222, RR 222, RY 145, R 6G
RR 189
AB 15, AB 25, AB 62, DR 81, MY 30, RB 5, RB 19
RR
AG 25, AO 10, AO 12, AR 18, AR 73
AB 40, AR 18, AR 88, DR 2
IC
azo dye
AO 7, AR 1, AR 88, AR 138, BB 9, CV
BB 1, BB 3
BB 9, CV, MO, O II
RR 141
AG 27
CV, O II, Rb 5, RB 5, RY 2
RR 195

An important class of chitosan derivatives are the


crosslinked materials, from gel types to bead types
or particles (including microparticles, microspheres
and nanoparticles). Gels are often divided into three
classes depending on the nature of their network,
namely entangled networks, covalently crosslinked
networks and networks formed by physical interactions. Berger et al. [26] suggested the following
modied classication for chitosan gels; i.e. a
separation of chemical and physical gels. Physical
gels are formed by various reversible links and
chemical gels are formed by irreversible covalent
links, as in covalently crosslinked chitosan gels.
Hydrogels and beads can be formed covalently
crosslinking polymer with itself. In this chemical
type of crosslinking reaction, the crosslinking agents
are molecules with at least two reactive functional
groups that allow the formation of bridges between polymer chains. To date, the most common

[71]
[72,73]
[74]
[75]
[76,77]
[78]
[7982]
[83]
[84,85]
[8693]
[94]
[95]
[96]
[9799]
[100]
[101]
[102]
[103105]
[106]
[107,108]
[109]
[110,111]
[112115]
[116]

crosslinkers used with chitosan are dialdehydes such


as glyoxal, formaldehyde and in particular glutaraldehyde (GLU) [26]. GLU reacts with chitosan and
it crosslinks in inter and intramolecular fashion
through the formation of covalent bonds mainly
with the amino groups of the polymer. Its reaction
with chitosan is very well documented. The main
drawback of GLU is that it is considered to be
toxic, even if the presence of free unreacted GLU in
gels is improbable since the materials are puried
before use. Other crosslinkers of chitosan are
epoxides such as epichlorohydrin (EPI) and ethylene glycol diglycidyl ether (EGDE), isocyanates
(hexamethylenediisocyanate) and other agents (carboxylic acids, genipin). Covalent crosslinking, and
therefore the crosslinking density, is inuenced by
various parameters, but mainly dominated by the
concentration of crosslinker. It is favoured when
chitosan molecular weight (MW) and temperature

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
Table 6
The four main classes of chitosan derivatives

Table 7
Some methods for preparation of chitosan particles

I. Modied polymers
 Carboxymethylchitosans
 Alkylated chitosans
 Chitosan sulfate derivatives
 Carbohydrate-branched chitosans
 Grafted chitosans
 Ligand-bound chitosan

Crosslinking with chemicals


 (Single) emulsion crosslinking
 Multiple emulsion
 Precipitation/crosslinking

II. Crosslinked chitosan


 Covalently crosslinked particles
 Ionically crosslinked particles
 Nanoparticles
 Physical gels
III. Chitosan-based composites
 Chitosan-dendrimer hybrids
 Chitosan-supported on inert materials (silica gel, glass beads,
alumina, etc.)
IV. Membranes

increased. Moreover, since crosslinking requires


mainly deacetylated reactive units, a high degree
of deacetylation (DD) of chitosan is favorable.
The crosslinked polymeric materials have a threedimensional network structure and can swell considerably in aqueous medium without dissolution.
Their synthesis and properties have been recently
described in detail [41]. Various methods have been
developed for the chemical crosslinking of chitosan,
which commonly result in gel formation. The
crosslinking step is a well-documented reaction
and an easy method to prepare chitosan-based
materials with relatively inexpensive chemicals.
Generally, a crosslinking step is required to
improve mechanical resistance and to reinforce the
chemical stability of the chitosan in acidic solutions,
modifying hydrophobicity and rendering it more
stable at drastic pH, which are important features to
dene a good adsorbent. However, it decreases the
number of free and available amino groups on the
chitosan backbone, and hence the possible ligand
density and the polymer reactivity. It also decreases
the accessibility to internal sites of the material and
leads to a loss in the exibility of the polymer
chains. So, the chemical step may cause a signicant
decrease in dye uptake efciency and adsorption
capacities, especially in the case of chemical reactions involving amine groups, since the amino
groups of the polymers are much more active
than the hydroxyl groups and can be much more
easily attacked by crosslinkers. Consequently, it is

411

Crosslinking and interactions with charged ions, molecules and


polymers
 Ionotropic gelation
 Wet-phase inversion
 Emulsication and ionotropic gelation
 Emulsication and solvent evaporation
 Simple or complex coacervation (precipitation, complexation)
Miscellaneous methods
 Thermal crosslinking
 Solvent evaporation method
 Neutralization method
 Spray drying
 Freeze drying
 Reverse micellar
 Coating
 Interfacial acylation

important to know, control and characterize the


conditions of the crosslinking reaction since they
determine and allow the modulation of the crosslinking density, which is the main parameter
inuencing interesting properties of gels [26]. These
conditions are useful for a better comprehension of
the adsorption mechanisms. For example, the loss
in exibility of the polymer resulting from the
crosslinking may explain some diffusion restrictions, and the decrease observed in the intraparticle
diffusivity.
Table 7 outlines various methods and approaches
which have been proposed for the preparation of
chitosan particles including microspheres/microparticles, and nanoparticles. Selection of any of the
methods depends upon factors such as particle
size requirement, thermal and chemical stability. In
practice, the methods are often combined and
different follow-up treatments are carried out [33].
The emulsion crosslinking method is widely used for
the synthesis of microspheres. This method is
schematically represented in Fig. 5. With this
method, the size of the particles can be controlled
by modifying the size of the aqueous droplets.
Another interesting method is spray drying. This is
a complex operation with the movement of countless droplets/particles in turbulent drying medium
ows under changing temperature and humidity

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

412

chitosan aqueous
solution

oil phase

emulsification

crosslinking agent
stirring

hardening of
droplets
separation
particles
Fig. 5. Schematic representation of preparation of chitosan
particles by emulsion crosslinking.

conditions. Chitosan microspheres obtained by this


technique are characterized by a high degree of
sphericity and specic surface area, parameters that
are important for application as adsorbents.
Ionic crosslinking reactions have also been
employed by using ionotropic gelation to form
hydrogels, beads and nanoparticles. Aside from its
complexation with negatively charged ions or
molecules, an interesting property of chitosan is its
ability to gel on contact with specic polyanions.
This gelation process is due to formation of inter
and intramolecular crosslinks mediated by these
polyanions. Tripolyphosphate (TPP) is commonly
used to provoke the ionotropic gelation of chitosan.
The particles can be obtained by the addition of a
chitosan solution to a solution of TPP or vice versa,
under strirring. In either case, the size of the
particles is strongly dependent on the concentration
of the solutions. Chiou and Li [68] and Szetos
group [110,111] recently reported the ionotropic
gelation of chitosan with TPP. They prepared
chitosan particles by adding an alkaline phase
containing TPP into an acidic phase containing
chitosan. (Nano)particles are formed immediately
upon mixing the two phases through molecular
linkages created between TPP phosphates and
chitosan amino groups. The solution of TPP was
used to produce more rigid materials. They reported
that TPP had no effect on dye adsorption. To

stabilize chitosan in acid solutions, Chiou and Li


[68] also proposed an ionotropic gelation process
followed by a chemical crosslinking step.
Chitosan is usually used in a aked or powdered
form that is both soluble in acidic media and nonporous. Moreover, the low internal surface area of
the non-porous polymer limits access to interior
adsorption sites and hence lowers dye adsorption
capacities and adsorption rates. To overcome this
obstacle, porous beads were synthesized. Indeed
an interesting characteristic of the chitosan is its
excellent ability to be processed into porous
structures.
3. A brief review of the recent literature on the
adsorption of dyes by chitosan
There is abundant literature concerning the
evaluation of adsorption performances of raw
chitosan, especially in terms of adsorption capacity
(amount of dye adsorbed) or uptake. In a batch
system, the determination of the dye uptake rate by
a chitosan-based material is often based on the
equilibrium state of the adsorption system. At least
100 dyes, mainly anionic dyes, have been so far
studied. Chitosan has an extremely high afnity for
many classes of dyes (Table 8). In particular, it has
demonstrated outstanding removal capacities for
anionic dyes such as acid, reactive and direct dyes.
This is due to its unique polycationic structure.
The effectiveness of chitosan for its ability to
interact with dyes has been studied by numerous
workers. Juang and co-workers [8993] demonstrated the usefulness of chitosan for the removal of
reactive dyes. They found that the maximum
adsorption capacities of chitosan for RR 222, RB
222 and RY 145 were 1653, 1009 and 885 mg/g,
respectively [90]. Annadurai [59,60] and Crini et al.
[72] also reported that chitosan may be a useful
adsorbent for the efuent of textile mills because of
its high adsorption capacity. Uzun and Guzel
[112115] noted that chitosan can be used in the
studies of dyestuff adsorption in comparison with
most other adsorbents. This polysaccharide showed
a higher capacity for adsorption of dyes than CAC
and other low-cost adsorbents, as reviewed by Crini
[6]. Kim and Cho [71] also indicated that the
amount of RB 5 adsorbed on chitosan beads is
much greater than on CAC. Similar conclusions
were reached by Lima et al. [58] for the BB 9
adsorption. McKays group [9799] recently published a series of papers on the ability of chitosan to

Table 8
Results of batch studies for various dyes using chitosan
Dye

Effective pretreatment of
chitosan

Particle size

protonation

450900mm
355500 mm

Crab shell

Sspa

Protonation

Powder (crab)
Wet bead
Dried bead
Powder (crab)

Flake
Bead (crab)
Swollen bead
Flake
Bead (lobster)

20
25

4h
24 h
4 days
24 h
5 days
24 h
5 days
24 h
3 days
3 days
5 days
24 h
24 h

Crosslinking
355500 mm

Crosslinking
355500 mm
355500 mm
Protonation
Grafting

450900 mm
25

Grafting

25
10
0.99

Crosslinking
0.206
60100 mesh
Crosslinking
Crosslinking
Crosslinking
Crosslinking
Crosslinking

Crosslinking
Crosslinking

Crosslinking
Crosslinking
Crosslinking
Crosslinking

0.24

2 mm
2.8 mm
11.41 mm
0.715 mm
25 mm
850 mm1 mm
850 mm1 mm
850 mm1 mm
2.32.5 mm
2.32.5 mm
2.52.7 mm
3.53.8 mm
2.32.5 mm

350

3
4
4
6
9.6
3

3
9.5
5.5
6
3
8.47
6
4
2
3
3
6
4

11.8
12.3
2
3
11
11
11
3
3
3
3
6

25
30
25
30
25
30
30
30
25
25
30
20
25
30
30
25
60
25
26
30
47.5
35
30
30
25
30
30
30
30
25
30
60
40
20
30
30
30
30
30

4h
40 min
3 days
3 days
40 min
24 h
3h
5h
5 days
24 h
2h
5 days
60200 min
5 days
2 days
5 days
5 days
4 days
4 days
24 h
5 days
24 h
24 h
24 h
5 days
5 days
5 days
5 days
5 days

Equilibrium
model

Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir

qmb

296
645.1
525
2103.6
1940
922.9
1954
973.3
656
494
1940
693.2
728.2
76
50
166.5
222
202
121.9

Kinetic model

2383
37.18

Reference

Diffusionc

[65]
[9799]
[81]
[110]
[67]
[9799]
[67]
[9799]
[86]
[86]
[67]
[9799]
[9799]
[83]
[65]
[73]
[86]
[86]
[72]
[59]
[58]
[75]
[67]
[60]
[63]
[66]
[62]
[67]
[79]
[71]
[66]
[90]
[90]
[89]
[74]
[67]
[109]
[109]
[109]
[94]
[68]
[94]
[94]
[94]

Lagergren

Ho and McKay
Lagergren
Ho and McKay
Lagergren
Elovich
Elovich
Ho and McKay
Lagergren
Lagergren
Ho and McKay
Ho and McKay
Elovich
Elovich
Lagergren

Langmuir
Langmuir
Langmuir

Adsorption
mechanism

Chemisorption
Chemisorption

Chemisorption
Diffusionc
Chemisorption
Chemisorption
Chemisorption
Chemisorption
Diffusionc

Lagergren
Ho and McKay

Langmuir
1334
Langmuir
Langmuir
Freundlich

2498
1100

Langmuir
Langmuir

722
1009
199

Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir

30.4
2422
156
110
68
1936
1834
1686
1642
1189

Ho and McKay
Avrami
Ho and McKay

Diffusionc

Ho and McKay
Ho and McKay
Ho and McKay

Diffusionc
Chemisorption
Chemisorption
Diffusionc

Ho and McKay

Ho
Ho
Ho
Ho
Ho

and
and
and
and
and

McKay
McKay
McKay
McKay
McKay

Chemisorption
Diffusionc
Chemisorption
Chemisorption
Chemisorption

413

Bead (crab)
Shrimp shell
Shrimp shell
Shrimp shell
Bead
Bead
Bead
Bead
Bead

3.6
4
3
4
4
3
4

355500 mm

Grafting

Shrimp shell
Bead (crab)
Bead
Bead (crab)

Equilibrium
time

180 nm
Crosslinking

0.177

Bead (crab)

T (1C)

ARTICLE IN PRESS

Nanoparticle
Bead (crab)
Crab shell
Bead (crab)
Crab shell
Wet bead
Dried bead
Bead (crab)
Crab shell
Crab shell
Bead (shrimp)

pH

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

AB
AG 25
AG 25
AG 27
AO 7
AO 10
AO 12
AO 12
AO 51
AO 51
AR 14
AR 18
AR 73
AR 87
BB
BB 3
BB 9
BB 9
BB 9
BB 9
BB 9
DB
DR 81
DS
IC
MY
RB
RB 2
RB 5
RB 5
RB 15
RB 222
RB 222
RB 222
RO 16
RR 2
RR 141
RR 141
RR 141
RR 189
RR 189
RR 189
RR 189
RR 189

Chitosan

414

Table 8 (continued )
Dye

2
86
145
145
145

Bead
Bead
Swollen bead
Wet bead
Dried bead
Bead (crab)
Bead (shrimp)
Bead (lobster)
Flake (shrimp)
Flake (lobster)
Flake
Flake (crab)
Bead
Bead (lobster)
Bead
Bead (crab)
Bead (crab)
Swollen bead
Flake
Bead (lobster)

Crosslinking

Sspa

2.32.5 mm

pH

T (1C)

Equilibrium
time

Equilibrium
model

qmb

Kinetic model

Adsorption
mechanism

Reference

6
3

30
30
30
30
30
30
30
30
30
30
30
30
30
30
30

5
2
5
4
3
3
5
5
5
5
5
4
5
3

Langmuir
Langmuir
Freundlich
Langmuir
Freundlich
Freundlich
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Langmuir
Freundlich

950
2252
1965
1653
1498
1215
1106
1026
1037
494
398
339
293

Ho and McKay
Ho and McKay
Lagergren
Ho and McKay
Elovich
Elovich

Diffusionc
Chemisorption
Diffusionc
Chemisorption
Chemisorption
Chemisorption
Diffusionc
Diffusionc
Diffusionc
Diffusionc
Diffusionc
Chemisorption
Diffusionc

[68]
[69]
[87]
[90]
[86]
[86]
[91]
[91]
[91]
[91]
[91]
[90]
[91]
[88]
[89]
[62]
[67]
[67]
[90]
[90]
[89]

2.8 mm

3.11 mm
2.39 mm
2.93 mm
1630 mesh
1630 mesh
11.41 mm
1630 mesh

3040
3040
3040
46
46
11.8
46

0.715 mm

12.3
0.24

Crosslinking

4.01

Crosslinking
Crosslinking
Crosslinking

Specic surface area in m2/g.


Adsorption capacities in mg/g.
c
Intraparticle diffusion model.
b

Particle size

2.8 mm
11.41 mm
0.715 mm

11.8
12.3

2
4
3

30
30
30
30
30

days
days
days
days
days
days
days
days
days
days
days
days
days
days

Ho and McKay

Diffusionc
60200 min
5 days
5 days
4 days
4 days

Langmuir
Langmuir
Langmuir
Langmuir

2436
1911
885
188

Avrami
Ho and
Ho and
Ho and
Ho and

McKay
McKay
McKay
McKay

Chemisorption
Chemisorption
Diffusionc

ARTICLE IN PRESS

189
222
222
222
222
222
222
222
222
222
222
222
222
222
222

Effective pretreatment of
chitosan

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RR
RY
RY
RY
RY
RY
RY

Chitosan

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

act as an effective adsorbent for the removal of acid


dyestuffs from aqueous solution. The monolayer
adsorption (saturation) capacities were determined
to be 973.3, 922.9, 728.2 and 693.2 mg of dye per
gram of chitosan for AO 12, AO 10, AR 73 and AR
18, respectively [99]. The interaction between
chitosan and anionic dyes has also been intensively
investigated by Guibal and co-workers [7982].
Their investigations clearly indicated that chitosan
had a natural selectivity for dye molecules and was
very useful for the treatment of wastewater. They
reported that adsorption capacities ranged between
200 and 2000 mmol/g for chitosan and between 50
and 900 mmol/g for CAC [82]. They concluded that
chitosan exhibited a twofold or more increase in the
adsorption capacity compared to CAC in the case of
acid, direct, reactive and mordant dyes. The best
choice for the adsorbent between CAC and chitosan
depends on the dye, however, it was impossible to
determine a correlation between the chemical
structure of the dye and its afnity for either carbon
or chitosan.
It is evident from this brief literature survey that
chitosan can be utilized as an interesting tool for the
purication of dye-containing wastewater because
of its outstanding adsorption capacity.
4. Control of adsorption performances of chitosan
The data from the literature show that the control
of adsorption performances of a chitosan-based
material in liquid-phase adsorption depends on the
following factors:
(i) the origin and nature of the chitosan such as
its physical structure, chemical nature and
functional groups;
(ii) the activation conditions of the raw polymer
(physical treatment, chemical modications);
(iii) the inuence of process variables such as
contact time, initial dye concentration, polymer
dosage and stirring rate;
(iv) the chemistry of the dye (e.g. its pKa, polarity,
MW size and functional groups);
(v) and nally, the solution conditions, referring to
its pH, ionic strength, temperature and presence
of impurities.
These aspects will be described in the following.
However, the reader is encouraged to refer to the
original papers for complete information on experimental conditions in the batch studies used.

415

4.1. Influence of the chitosan characteristics


It is very important to note that tuning the chitosan
manufacturing process can ernable the production of
polymers with varying chemical characteristics and
MW distributions. As stated in the introduction,
chitosan is a collective term applied to deacetylated
chitins in various stages of deacetylation and
depolymerization [37]. Commercial chitosan is usually offered as akes or powders. Products of various
companies differ in purity, salt-form, color, granulation, water content, DD or degree of acetylation
(DA), amino group content, MW, crystallinity and
solubility [1012,18]. These parameters determined by
the conditions selected during the preparation are
very important because they control the swelling and
diffusion properties of chitosan and also inuence its
characteristics [117]. In particular, numerous studies
have demonstrated that the MW and DD inuence
the adsorption properties of this polymer. Therefore,
these factors must be considered carefully during the
adsorption optimization process.
4.1.1. Chitosan origin
From a practical viewpoint, crustaceans shells are
the potential sources for chitin production. Chitosan is commonly prepared by deacetylating chitin
using 4050% aqueous alkali at 110115 1C for a
few hours [12]. Chitin occurs in a wide variety of
species, from fungi to animals. Depending on the
chitin source, chitosan varies greatly in its adsorption properties and solution behavior, as reported
by Juang and co-workers [8993]. For example, the
adsorption capacities of RR 222 on different types
of chitosan prepared from three shery wastes
(shrimp, crab and lobster shells) were compared.
The monolayer adsorption capacities were determined to be 293, 398 and 494 mg of dye per gram of
ake-type of chitosan for crab, lobster and shrimp,
respectively [91]. This demonstrates that the adsorption capacity of chitosan depends on its origin.
Rinaudo [11] also reported in a recent review that
the origin of chitin inuences not only its crystallinity and purity but also its polymer chains
arrangement, and hance its properties. In particular,
the chitin resulting from crustaceans needs to be
graded in terms of purity and color since residual
protein and pigment can cause problems [10,11].
4.1.2. Physical nature of the chitosan
The adsorption capacity of chitosan also depends
on its physical structural parameters such as

ARTICLE IN PRESS
416

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

crytallinity, surface area, porosity, particle type,


particle size and water content. These parameters
are determined by the conditions selected during the
preparation and polymer conditioning.
Three crystalline forms are known for chitin:
a-, b- and g-chitins. The most abundant and easily
accessible form is a-chitin [11,91]. Chitosan is also
crystalline and shows polymorphism depending on
its physical state. Depending on the origin of the
polymer and its treatment during extraction from
raw resources, the residual crystallinity may vary
considerably. Crystallinity is maximum for both
chitin (i.e. 0% deacetylated) and fully deacetylated
chitosan (i.e. 100%). Generally, commercial chitosans are semi-crystalline polymers and the degree
of crystallinity is a function of the DD. Crystallinity
plays an important role in adsorption efciency as
reported by Trung et al. [108]. They demonstrated
that decrystallized chitosan is much more effective
in the adsorption of anionic dyes. Crystallinity
controls polymer hydratation, which in turn determines the accessibility to internal sites. This parameter strongly inuences the kinetics of hydratation
and adsorption. Dissolving the polymer breaks the
hydrogen bonds between polymer chains. The
reduced polymer crystallinity can be maintained
through freeze-drying of the chitosan solution,
while air-drying or oven-drying partially reestablishes polymer crystallinity. The conditioning of the
polymer and physical modication can strongly
reduce the inuence of this important parameter
and improve diffusion properties [18]. The gel
formation procedure also allows an expansion of
the polymeric network, a decrease in steric hindrance phenomena and a decrease in the crystallinity of raw materials which enhance mass
transport. The case of dye adsorption with crosslinked chitosan is a typical example of the inuence
of particle size. When crosslinked with GLU, the
network formed makes the sorption performances
become dependent on the size of particles. This
dependence disappears when chitosan particles are
modied by gel formation. Hebeish et al. [84,85]
indicated that the crosslinking step changes the
crystalline nature of chitosan and decrease the
particle size of the crystallites, enhancing its
adsorption capacity. The crosslinking reaction
destroys the crystalline structure at low levels of
crosslinking. The authors assumed that more
accessible domains are created as a result of changes
in the physical and chemical structures of chitosan
during the modication by GLU, and consequently

these effects increased dye adsorption [85]. However, Cestari et al. [62] recently noted that after the
crosslinking reaction, there is a small increase in the
crytallinity of chitosan beads with increased access
to the small pores of the material.
Among the other parameters that have a great
impact on dye adsorption is particle type. Chitosan
can be presented as gels, akes, powders and
particles. Chitosan beads are preferred since ake
and powder forms of polymer are not suitable for
use as adsorbents due to their low surface area and
lack of porosity, as indicated by Varma et al. [19].
Beads are usually prepared by dropping highviscosity chitosan salt solutions into a basic solution
with slow stirring. The diameters of the drops as
well as the solution ow rate control the diameter of
the beads. Wu et al. [91] reported that bead-type
chitosan gives a higher capacity for dye adsorption
than the ake type by a factor of 24 depending on
the source of shery waste. For example, a
comparison of the maximum adsorption capacity
(qmax) for RR 222 by chitosan akes and beads
prepared from a crab source showed 293 mg/g for
akes and 1103 mg/g for beads. The authors
explained this result by the fact that the beads
possessed a greater surface area (i.e., more loose
pore structure) than the akes. They also reported
that the adsorption capacity of chitosan depends on
its source. The qmax were determined to be 1106,
1037 and 1026 mg of dye per gram of bead-type of
chitosan for crab, lobster and shrimp, respectively
[91]. Again, it can be noted that the order of qmax for
the different sources is exactly identical to that of
the surface area of the whole animal, i.e., crab4
lobster4shrimp. Chang and Juang [86] also noted
that chitosan in the bead form signicantly improves the adsorption performance of RR 222, AO
51 and BB 9 compared to that in the ake form.
Guibal et al. [82] indicated that it would be
interesting to use chitosan gel beads instead of
akes since the production of gel beads decreases
the residual crystallinity of polymer which enhances
both the porosity and the diffusion properties of the
material, due to the expansion of the chitosan
network and the increase in the specic surface area.
Crini et al. [72] observed that compared to chitosan
akes, chitosan beads exhibited a twofold or more
increase in the adsorption capacity for BB 9. One of
chitosans most promising features is its excellent
ability to be processed into nanostructures. These
nanochitosans can also be used in batch studies, as
reported by Hu et al. [110]. They noted that an

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

adsorption capacity of 2103.6 mg of AG 27 per


gram chitosan was achieved, which was signicantly
higher than that of the chitosan microparticles.
Previously, it has been demonstrated that the
particle size of chitosan also inuences its adsorption prole. For example, Park et al. [56] showed
that of the smaller particle size, the more dye was
absorbed. As adsorption is a surface phenomenon,
this can be attributed to the relationship between
the effective specic surface area of the adsorbent
particles and their sizes. The surface area values
usually increased as the particle size decreased and,
as a consequence, the saturation capacity per unit
mass of adsorbent increased. Decreasing the size of
particles improves the adsorption properties of the
chitosan, especially when chitosan is crosslinked.
However, small particle sizes are not compatible
with large-scale applications. For example, in xedbed columns, small particles are inappropriate since
they induce head loss and column blocking and
cause serious hydrodynamic limitations [32]. There
are a large number of studies that highlight the
correlation between adsorption performance and
size of particle. Annadurai [59,60] used chitosan for
the removal of basic and direct dye from solutions.
The results indicated that the adsorption efciency
depends upon the particle size, dosage and temperature. In particular, the adsorption capacity
increased with a decrease in the particle size and
the dye molecules were preferably adsorbed on the
outer chitosan surface. The author suggested that
this observation can be attributed to the larger total
surface associated with smaller particles [60]. In
contrast to the ndings of Annadurai, Guibal and
co-workers [8082] observed that the adsorption
occurred not only at the surface of the material
due to rapid surface adsorption but also in the
intraparticle network of the polymer. In particular,
the large external surface area for small particles
removes more dye in the initial stages of the
adsorption process than the large particles, conrming the previous results reported by McKay
et al. [44,45]. They studied the adsorption of AG 25
on chitosan and reported that the size of adsorbent
particles inuenced both the adsorption kinetics and
equilibrium [81] because of the resistance to
intraparticle diffusion. The greater the particle size,
the greater the contribution of intraparticle diffusion resistance to the control of the adsorption
kinetics for materials of low porosity. In other
works [80,82], they indicated that the time required
to reach equilibrium increased on increasing the size

417

of the adsorbent particles. This means that intraparticle diffusion greatly inuences the accessibility
of dye molecules to internal sites. With raw
chitosan, the differences were more marked than
with protonated material [80]. Due to resistance to
intraparticle mass transfer in raw chitosan, it is
usually necessary to use very small particles to
improve adsorption kinetics. When the dyes have
strong interactions with chitosan, this allows larger
adsorbent particle sizes to be used to get the same
adsorption rate. They concluded that this was
especially interesting for large-scale applications
since it was easier to manage large adsorbent
particles rather than ne powders [82]. Juang et al.
[93] also observed that the adsorption capacity
strongly depended on the particle size of chitosan.
At a chitosan particle size of 250420 mm, the values
were 380, 179 and 87 mg/g for RR 222, RY 145 and
RB 222, respectively. These results were signicantly greater than those obtained using adsorbents
such as CAC, natural clay, bagasse pith and maize
cob, in which the capacity for reactive dyes was
often less than 30 mg/g. They concluded than the
smaller the chitosan particles, the greater the
capacity for dye. Li and co-worker [94] reported
similar conclusions for the adsorption of basic dyes
on the adsorption of RR 189 on crosslinked beads.
For example, the adsorption capacity of particles
with diameters 2.32.5, 2.52.7 and 3.53.8 were
1936, 1686 and 1642 mg/g, respectively, at pH 3 and
30 1C. They also concluded that the dye uptake
increased with a decrease in the particle size
since the effective surface area was higher for the
same mass of smaller particles. Chiou and Chuang
[66], using crosslinked chitosan for the removal of
dye from solutions, indicated that the increase in
adsorption capacity with decreasing particle size
suggests that the dye preferentially adsorbed on the
outer surface and did not fully penetrate the particle
due to steric hindrance of large dye molecules.
Recently, Trung et al. [108] reported that no effect
of the difference in particle size of decrystallized
chitosan on the decolorization capacity was observed. The size of particles has been shown to be a
key parameter in the control of adsorption performances of several dyes on chitosan, in particular
this may be the main parameter to control dye
adsorption equilibrium. However, the relationship
of adsorption capacity to particle size also principally depends on two criteria: (i) the chemical
structure of the dye molecule (its ionic charge) and
its chemistry (its ability to form hydrolyzed species)

ARTICLE IN PRESS
418

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

and (ii) the intrinsic characteristic of the adsorbent


(its crystallinity and porosity, the rigidity of the
polymeric chains, the degree of crosslinking), as
shown by Guibal and co-workers [8082].
Adsorption performance (in particular intraparticle diffusion) is also controlled by polymer
porosity (i.e. porous volume, porous distribution
and pore size). CAC are well-known conventional
porous adsorbents and are characterized by a large
specic surface area and a great porosity that limits
the resistance to intraparticle diffusion. The aggregation of dye molecules may involve a strong
increase in the size of the diffusing molecule, and
this effect may be reinforced by the inuence of pore
size in controlling intraparticle diffusion properties
and accessibility to internal sites. Thus, the efciency in adsorbing dyes onto a material such as
CAC can be correlated to its surface characteristics.
However, chitosan is known as a non-porous
polymer. It is characterized by a low surface area
and a low porosity that control the diffusion to the
center of the particles, especially with large molecules. These features generally limit access to
interior adsorption sites. So, polymer porosity may
affect the dye adsorption capacity of chitosan. In
crosslinked chitosan beads, usually prepared by a
chemical treatment with GLU, the materials are
submicron to micron-sized, and need large internal
pores to ensure adequate surface area for adsorption. Indeed these chemical treatments involve
supplementary linkages that limit the transfer of
solute molecules. In general, diffusion limitation
within particles leads to the decreases in adsorption.
These limiting effects can be compensated for by the
physical modication of the polymer. As already
mentioned, an interesting characteristic of chitosan
is its excellent ability to be processed into porous
and nanoporous structures. Gel bead conditioning
in addition to the decrease of polymer crystallinity,
improves both swelling and diffusing properties, but
also allows expansion of the porous structure of the
network, which in turn enhances the transport
of dyes. This physical modication allows both
the polymer network to be expanded (enhancing
the diffusion of large sized molecules) and the
crystallinity of the polymer to be reduced. Porous
structures can be formed by freeze-drying chitosanacetic acid solutions in suitable molds. Exclusion of
chitosan acetate salt from the ice crystal phase and
subsequent ice removal by lyophilization generates
a porous material with a mean pore size that can be
controlled by varying the freezing rate and hence the

ice crystal size. Pore orientation can be directed by


controlling the geometry of thermal gradients
during freezing. The mechanical properties of the
resulting material are mainly dependent on the pore
sizes and pore orientations. Another process consists in dissolving the polymer in acid solution
followed by a coagulation. Recently, Kim and Cho
[71] proposed a solgel method to prepare porous
chitosan beads with interesting high internal specic
surface areas, allowing better accessibility of dyes to
interior adsorption sites. Nanotechnology has been
also proposed to prepare porous materials
[110,118,119]. Compared to the traditional micronsized materials, nano-sized adsorbents possess quite
good performance due to high specic area and
porous structure, and the absence of internal
diffusion resistance.
4.1.3. Chemical structure of chitosan
The properties of chitosan also depend on its
chemical nature (MW, DD), functional groups
(ionic charge, variety, density, accessibility) and
solution behavior (purity, water content, salt-form,
afnity for water). These parameters are also
determined by the conditions selected during the
preparation.
It is known that chitin samples have different DD
depending on their origin and mode of isolation
[12]. Deacetylation takes place during isolation by
alkaline treatment to remove proteins. To prepare
chitin with a fully N-acetylated polymer or a
uniform structure, selective N-acetylation of the
free amino groups is necessary. Chitosan is prepared
by deacetylating chitin. Depending on the chitin
source and the methods of hydrolysis, commercial
chitosan also varies greatly in its MW and distribution, and therefore its solution behavior. The MW
of chitosan is a key variable in adsorption properties
because it inuences the polymers solubility and
viscosity in solution. It is an important factor for
characterization, but poor solubility and structural
ambiguities in connection with the distribution of
acetyl groups are major obstacles to quantitatively
determining MWs [11]. It is also difcult to
determine the MW of native chitin.
Another important characteristic of chitosan is
the degree of N-acetylation (DA) or DD. The DD
parameter is essential since, though the hydroxyl
groups on the polymer may be involved in attracting
dye molecules, the amine functions remain the main
active groups and so can inuence the polymers
performance. Guibal et al. [82] observed that

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

increasing the DD involved an increase in the


relative proportion of amine groups, which were
able to be protonated, favoring dye adsorption.
However, they indicate that the variation in
adsorption properties was not proportional to
DD, but changed with the type of dye, especially
with chitin. Saha et al. [102], studying the adsorption of an azo dye onto chitosan akes, also
reported that the results were found to be strongly
dependent on the DD of the polymer. The higher
DD chitosan provided a better adsorption. Recently, it has been reported that the solution
properties of a chitosan depend not only on its
average DA but also on the distribution of the
acetyl groups along the main chain [11]. However,
Chiou and Li [68], studying the adsorption
of RR 189 on crosslinked chitosans reported
that both the MW and the DD of the polymer
were almost without effect on the adsorption
capacities.
An additional advantage of chitosan is the high
hydrophilic character of the polymer due to the
large number of hydroxyl groups present on its
backbone. Depending on its MW and DD, chitosan
in aqueous solution is expected to have the properties of an amphiphilic polymer. With an increase in
DD, the number of amino groups in the polymer
increases, and with an increase of MW, the polymer
conguration in solution becomes a chain or a ball.
In addition, adsorption is known to change the
conformation of the chitosan polymer. The viscosity
of chitosan also greatly inuences the chitosan
conditioning processes.
4.2. Activation conditions
4.2.1. Chitosan preprotonation
Because of its stable, crystalline structure, the
polyamine chitosan is insoluble in either water or
organic solvents. However, in dilute aqueous acids,
the free amino groups are protonated and the
polymer becomes fully soluble below pH 5. Since
the pKa of the amino group of glucosamine residues
is about 6.3, chitosan is extremely positively charged
in acidic medium. So, treatment of chitosan with
acid produces protonated amine groups along the
chain and this facilitates electrostatic interaction
between polymer chains and the negatively charged
anionic dyes, as previously observed by Maghami
and Roberts [50]. The pH-dependent solubility of
chitosan provides a convenient tool to improve its
performance although solubility is a very difcult

419

parameter to control [11]. In fact, the solubility and


its extent depends on the concentration and on the
type of acid. The polymer dissolves in hydrochloric
acid and organic acids such as formic, acetic, lactic
and oxalic acids. However, solubility decreases with
increasing concentrations of acid. Solubility is also
related to the DA, the ionic concentration, as well as
the conditions of isolation and drying of the
polymer [11]. In particular, the distribution of acetyl
groups along the chain (random or blockwise) can
strongly inuence the solubility of the polysaccharide and also the interchain interactions due to
H-bonds and the hydrophobic character of the
acetyl groups.
Trung et al. [108] proposed a pretreatment using
citric acid to produce decrystallized chitosan with a
low degree of crystallinity and a high anionic dyebinding capacity. The percentage crystallinity of
decrystallized chitosan was 10%, signicantly lower
than that of raw chitosan (32%). This reduction is
attributed to a probable rearrangement of polymer
chains during precipitation in the presence of citrate
ions. They also indicated that the decrystallized
chitosan had the same degree of DD and MW as the
original chitosan. Decrystallized chitosan adsorbed
anionic dyes almost twofold more efciently than
raw chitosan, due to its more amorphous character,
but showed decreased adsorption for cationic dyes.
However, the presence of ash in decrystallized
chitosan could also play a role in increased dyebinding capacity. Gibbs et al. [81] showed that the
preliminary protonation of amine groups, obtained
by contact with a sulfuric acid solution, reduced
the variation of solution pH following adsorbent
addition. Crini et al. [72] found that a homogeneous
chemical treatment such as a solubilization
reprecipitation process could give a chitosan product with a higher adsorption level for dyes than
one prepared by a heterogeneous process with the
same DD. They attributed this to an increase of the
surface area due to the conversion of the chitosan
akes into a powder.
4.2.2. Grafting reactions
Several workers have suggested that although
chitosan as such is very useful for treating contaminated solutions, it may be advantageous to
chemically modify chitosans, e.g. by grafting reactions [72,73,77,95,103,106,113]. The modications
can improve chitosans removal performance and
selectivity for dyes, alter the physical and mechanical properties of the polymer, control its diffusion

ARTICLE IN PRESS
420

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

properties and decrease the sensitivity of adsorption


to environmental conditions. Chemical grafting
of chitosan with specic ligands has been reviewed
by Jayakumar et al. [23] and by Prabaharan and
Mano [41].
It is known that the only class for which chitosan
[106] and crosslinked chitosan [85] have low afnity
is basic (cationic) dyes. To overcome this problem,
Crini et al. [72,73] suggested the use of N-benzyl
mono- and disulfonate derivatives of chitosan in
order to enhance its cationic dye hydrophobic
adsorbent properties and to improve its selectivity.
The maximum adsorption capacities of these
adsorbents for BB 9 and BB 3 were 121.9 and
166.5 mg/g, respectively. These derivatives could be
used as hydrophobic adsorbents in acidic media
without any crosslinking reaction. To fully develop
the high potentials of chitosan, it is necessary to
introduce chemical substituents at a specic position
in a controlled manner as suggested by Lima et al.
[58] and Chao et al. [106]. Lima et al. [58] proposed
the use of chitosan chemically modied with
succinic anhydride in the BB 9 adsorption. This
chemical derivatization provides a powerful means
to promote new adsorption properties in particular
towards basic dyes in acidic medium. Chao et al.
[106] suggested enzymatic grafting of carboxyl
groups onto chitosan as a means to confer the
ability to adsorb basic dyes on beads. The presence
of new functional groups on the surface of beads
results in increases in surface polarity and the
density of adsorption sites, and this improves the
adsorption selectivity for the target dye. Other
studies showed that the ability of chitosan to
selectively adsorb dyes could be further improved
by chemical derivatization. Shimizu et al. [103]
proposed novel chitosan-based materials by reacting chitosan with a higher fatty acid functionalized
with a glycidyl moiety in order to introduce long
aliphatic chains. They observed that these products
could be used as effective adsorption materials for
both anionic and cationic dyes. Martel et al. [95],
and El-Tahlawy and co-workers [76,77] proposed
the use of cyclodextrin-grafted chitosan as new
chitosan derivatives for the removal of dyes. Martel
et al. [95] demonstrated that these materials are
characterized by a rate of adsorption and a global
efciency greater than that of the parent chitosan
polymer. Uzun and Guzel [113,114] reported that
carboxymethylated chitosan is a rather better
adsorbent than raw chitosan for acidic dyestuffs,
and its production is not costly.

4.2.3. Influence of crosslinking


Raw chitosan powders also tend to present some
disadvantages such as unsatisfactory mechanical
properties and poor heat resistance. Another
important limitation of the raw material is that it
is soluble in acidic media and therefore cannot be
used as an insoluble adsorbent under these conditions, except after physical and chemical modication. One method to overcome these problems is to
transform the raw polymer into a form whose
physical characteristics are more attractive. So,
crosslinked beads have been developed and proposed. After crosslinking, these materials maintain
their properties and original characteristics [62],
particularly their high adsorption capacity,
although this chemical modication results in a
decrease in the density of free amine groups at the
surface of the adsorbent in turn lowering polymer
reactivity towards metal ions [80].
An important work on crosslinked chitosan was
done by Chiou and co-workers [6670]. Chitosan
beads were crosslinked with GLU, EPI or EGDE.
The results showed that the chitosan-EPI beads
presented a higher adsorption capacity than GLU
and EGDE resins [68,69]. They reported that these
materials can be used for the removal of reactive,
direct and acid dyes. It was found that 1 g chitosan
adsorbed 2498, 2422, 2383 and 1954 mg of RB 2,
RR 2, DR 81 and AO 12, respectively [67]. It is
important to specify that the adsorption capacities
of CAC for reactive dyes generally vary from 278 to
714 mg/g [6]. Another advantage of EPI is that it
does not eliminate the cationic amine function of
the polymer, which is the major adsorption site to
attract the anionic dyes during adsorption [69]. The
crosslinking of chitosan with GLU (formation of
imine functions) or with EDGE decreases the
availability of amine functions for the complexation
of dyes and with a high crosslinking ratio the uptake
capacity drastically decreases. They also indicated
that the crosslinking ratio slightly affected the
equilibrium adsorption capacity for the three cross
linkers under the range they studied [68]. The
amount of dye adsorbed was found to be higher in
acidic than in basic solution. This was explained by
considering the rate of diffusion from the swollen
beads in acidic and basic media. In basic medium, a
limited swelling of the beads inhibited the diffusion
of dyes at a faster rate as it occurred in acidic
medium. Among the conditions of the crosslinking
reaction that have a great impact on dye adsorption
are the chemical nature of the crosslinker, as

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

mentioned above, but also the extent of the


reaction. In general, the adsorption capacity depends on the extent of crosslinking and decreases
with an increase in crosslinking density. When
chitosan beads were crosslinked with GLU under
heterogeneous conditions, it was found that the
saturation adsorption capacity of RR 2 on crosslinked chitosan decreased exponentially from 200 to
50 mg/g as the extent of crosslinking increased from
0 to 1.6 mol GLU/mol of amine. This is because of
the restricted diffusion of molecules through the
polymer network and reduced polymer chain
exibility. Also the loss of amino-binding sites by
reaction with aldehyde is another major factor in
this decrease. However, Chiou and co-workers
indicated that the crosslinking step was necessary
to improve mechanical resistance, to enhance the
resistance of material against acid, alkali and
chemicals, and also to increase the adsorption
abilities of chitosan. The removal performance of
crosslinked chitosan and CAC for anionic dyes were
compared: the adsorption values were 315 times
higher at the same pH. Chiou and co-workers
[6670] concluded that chitosan chelation was the
procedure of choice for dye removal from aqueous
solution. However, Kim and Cho [71], studying the
adsorption of RB 5 on crosslinked chitosan beads,
arrived at contrasting conclusions. They demonstrated that the adsorption capacity of non-crosslinked beads was greater than that of crosslinked
beads in the same experimental conditions.
The materials, mainly crosslinked using GLU,
have been also proposed as effective dye removers
by several other workers [62,77,84,85,88,94,105]. All
these studies showed that the reaction of chitosan
with GLU leads to the formation of imine groups,
in turn leading to a decrease in the number of amine
groups, resulting in a lowered adsorption capacity,
especially for dyes sorbed through ion-exchange
mechanisms. However, this limiting effect of a
chemical reaction with GLU signicantly depends
on both the procedure used and the extent of
crosslinking, as reported by Hebeish et al. [84,85]. In
heterogeneous conditions, chitosan (solid state) was
simply mixed with GLU solution, while in homogeneous conditions chitosan was mixed with GLU
solution after being dissolved in acetic acid solution.
An optimum aldehyde/amine ratio was found for
dye adsorption, which depended on the crosslinking
operation mode (water-soluble or solid-state solution). The initial increase in dye adsorption was
attributed to the low levels of crosslinking in the

421

precipitates preventing the formation of closely


packed chain arrangements without any great
reduction in the swelling capacity. This increase in
adsorption was interpreted in terms of the increases
in hydrophilicity and accessibility of complexing
groups as a result of partial destruction of the
crystalline structure of the polymer by crosslinking
under homogeneous conditions. At higher levels of
crosslinking, the precipitates had lower swelling
capacities, and hence lower accessibility because of
the more extensive three-dimensional network and
also because of its more hydrophobic character with
increased GLU content. Juang et al. [88], studying
the adsorption of RR 222 on crosslinked chitosan
beads, also observed that the adsorption capacity
depends on the extent of crosslinking and decreases
as crosslinking density increases. This result was
mainly interpreted by the fact that the crosslinking
reaction with GLU decreases the availability of
amine functions for the complexation of dyes. The
results showed that the chitosan-GLU beads presented a higher adsorption capacity than glyoxal
beads. Gaffar et al. [77] and Shimizu et al. [105]
reported that the extent of crosslinking showed a
signicant inuence on adsorption properties. These
authors noted that the increase in the extent of
crosslinking is accompanied by a decrease in dye
uptake, conrming the results of Hebeish et al.
[84,85]. The adsorption capacity increased greatly at
low degrees of substitution but decreased with
increasing substitution. This phenomenon is interpreted in terms of increased hydrophilicity caused
by the destruction of the crystalline structure at low
crosslinking densities, while this can be associated
with an accompanying decrease in active sites,
accessibility, and swellability of the adsorbent by
increasing the level of crosslinking. On the contrary,
Chiou and Li [94], studying the adsorption of RR
189 on EPI-crosslinked chitosan beads, reported
that the crosslinking ratio did not affect the
adsorption capacity.
Another study showed that the physical and
mechanical properties of chitosan could be further
improved by crosslinking. Chitosan forms gels
below pH 5.5 and acid efuents could severely limit
its use as an adsorbent in removing dyes from acid
efuent. To solve this problem, Cestari et al. [62]
proposed the use of homogeneously crosslinked
beads. They reported that the beads were not only
insoluble in acid solution but also presented higher
specic surface areas (0.1 and 0.24 m2/g before and
after the crosslinking reaction, respectively) and

ARTICLE IN PRESS
422

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

stronger mechanical resistance than the raw chitosan powder. The chemical, physical and mechanical behavior of the beads and also adsorption
properties were enhanced by crosslinking with
functional groups. The materials had a strong
adsorption capacity for RY, RB and RR below
pH 5.5. The authors also noted that crosslinking can
change the crystalline nature of chitosan, as
suggested by the XRD diffractograms. After the
crosslinking reaction, there was a small increase in
the crytallinity of the chitosan beads and also
increased accessability to the small pores of the
material.
4.2.4. Chitosan-based composite beads
Practical industrial applications of raw chitosan
in xed-bed systems or packed in adsorption
columns are also limited. The characteristics of the
polymers can introduce hydrodynamic limitations
and column fouling, which limits their use for largescale columns. For example, the aked or powdered
form swells (the crosslinked beads have lower
swelling percentage [120]) and crumbles easily, and
does not function ideally in packed-column congurations common to pump-and-treat adsorption
processes. Various chitosan-based composites have
been designed to overcome these problems. Chang
and Juang [87] proposed the addition of activated
clay to chitosan to prepare composite beads in order
to improve its mechanical properties. Cestari et al.
[61] also proposed the use of silica/chitosan hybrid
for the removal of anionic dyes from aqueous
solutions: these materials are of interest because
they combine the structure, strength and chemical
properties of the silica with the specic characteristics of chitosan. Chang and Chen [64] proposed
the use of chitosan-conjugated Fe3O4 nanoparticles
for the removal acid dyes from aqueous solutions.
The adsorption capacities were 1883 and 1471 mg of
dye/g of chitosan for AO 12 and AG 25. Paneva
et al. [96] also proposed a novel effective route for
incorporating magnetic material into chitosan beads
by capillary extrusion. They concluded that the
material might be used for wastewater treatment in
the textile industry.
4.3. Influence of process variables
The amount of dye that can be removed from a
solution by chitosan also depends on process
variables used in batch systems such as chitosan

dosage, initial dye concentration, contact time,


agitation rate and dryness.
4.3.1. Effect of chitosan dosage
Of all the above factors, chitosan dosage is
particularly important because it determines the
extent of decolorization and may also be used to
predict the cost of chitosan per unit of solution to be
treated. As expected, the adsorption density increases signicantly as adsorbent dosage decreases.
This is due to the higher amount of the dye per unit
weight of adsorbent. Wen et al. [116] showed that
the increasing chitosan dose had a dramatic positive
impact on color removal and there was an
approximately linear relationship between chitosan
dose and color removal of the dye. Crini et al.
[72,73] also observed that the increase in adsorption
with adsorbent dosage can be attributed to increased adsorbent surface and availability of more
adsorption sites. However, if the adsorption capacity was expressed in mg adsorbed per gram of
material, the capacity decreased with the increasing
amount of sorbent. This may be attributed to
overlapping or aggregation of adsorption sites
resulting in a decrease in total adsorbent surface
area available to the dye and an increase in diffusion
path length. It was also indicated that the time
required to reach equilibrium decreased at higher
doses of adsorbent.
4.3.2. Effect of initial dye concentration
Park et al. [56] and Knorr [121] previously found
signicant correlations between dye concentration
and the dye-binding capacity of chitin or chitosan.
The amount of the dye adsorbed onto chitosan
increased with an increase in the initial concentration of dye solution if the amount of adsorbent was
kept unchanged. This is due to the increase in the
driving force of the concentration gradient with the
higher initial dye concentration. In most cases, at
low initial concentration the adsorption of dyes by
chitosan is very intense and reaches equilibrium
very quickly. This indicates the possibility of the
formation of monolayer coverage of the molecules
at the outer interface of the chitosan. At a xed
adsorbent dose, the amount adsorbed increased
with increasing concentration of solution, but the
percentage of adsorption decreased. In other words,
the residual concentration of dye molecules will be
higher for higher initial dye concentrations. In the
case of lower concentrations, the ratio of initial
number of dye moles to the available adsorption

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

sites is low and subsequently the fractional adsorption becomes independent of initial concentration
[67,68,75,83]. At higher concentrations, however,
the number of available adsorption sites becomes
lower and subsequently the removal of dyes depends
on the initial concentration. At the high concentrations, it is not likely that dyes are only adsorbed in a
monolayer at the outer interface of chitosan. As a
matter of fact, the diffusion of exchanging molecules within chitosan particles may govern the
adsorption rate at higher initial concentrations.
Recently, Gaffar et al. [77] reported that the
adsorption percentage decreases on increasing
the dye concentration. This could be ascribed to
the accompanying increase in dye aggregation
and/or depletion of accessible active sites on the
material.
4.3.3. Effect of contact time
Contact time is another important variable in
adsorption processes. Generally speaking, the adsorption capacity and the removal efciency of dyes
by chitosan become higher on prolonging the
contact time. However, in practice, it is necessary
to optimize the contact time, considering the
efciency of desorption and regeneration of the
adsorbent.
During the process, the adsorbent surface is
progressively blocked by the adsorbate molecules,
becoming covered after some time. When this
happens, the adsorbent cannot adsorb any more
dye molecules. As each particle puries a certain
volume of liquid, increasing the dosages rapidly
promotes an equilibrium between adsorbate and
adsorbent because the number of particles to treat
the same volume of liquid is increased. In general,
the adsorption capacity increases with time and, at
some point in time, reaches a constant value where
no more dye is removed from the solution. At this
point, the amount of dye being adsorbed onto the
material is in a state of dynamic equilibrium with
the amount of dye desorbed from the adsorbent.
The time required to attain this state of equilibrium
was termed the equilibrium time (te) and the amount
of dye adsorbed at te reected the maximum dye
adsorption capacity of the adsorbent under these
conditions.
Chatterjee et al. [83] reported an equilibrium time
of 20 h for eosin adsorption onto chitosan. They
observed that the process was initially very fast and
then slowly reached equilibrium. Dutta et al. [75]
noted that the maximum accumulation occurred

423

within 45 h for reactive and direct dyes on


chitosan. Guibal et al. [82], studying the adsorption
of 12 anionic dyes on chitosan also observed that
equilibrium was reached within the rst 12 h of
contact and adsorption kinetics were relatively fast.
Gibbs et al. [81] noted that with increasing AG 25
concentration relative to a xed adsorbent dosage,
the time required to reach equilibrium strongly
increased. Although 12 h was sufcient to achieve
complete recovery of the dye at initial concentrations of below 100 mg/l. For the highest concentration (200 mg/l) with raw chitosan 8 h was necessary
to reach equilibrium and the complete elimination
of the dye. Fahmy et al. [79] reported an equilibrium
time of 45 min for anionic dye adsorption on
crosslinked chitosan. Cestari et al. [61,62] found
reaction times of 60200 min. At the other extreme,
Wu et al. [91] observed that it took several days for
equilibrium to be attained. Crini et al. [72,73],
studying adsorption of basic dyes (BB 9 and BB 3)
with different chitosan concentrations and contact
times observed that 40 min of contact time was
enough to reach adsorption equilibrium in all the
experiments, while Chang and Juang [86] report that
3 days were required for BB 9 adsorption onto
chitosan. The difference might have been due to
differences in the properties of the material used or
to differences in the aqueous solution treated. Crini
et al. [73] used grafted chitosan with a solution
containing sodium chloride at pH 3 while Chang
and Juang [86] used chitosan that had not been
pretreated with a neutral solution. However, most
authors seem to agree on a gure in the range
35 days for most dye molecules (Table 8) and with
the fact that the adsorption of dyes is fast at the
initial stages of the treatment time, and thereafter,
becomes slower near the equilibrium. It is obvious
that a large number of vacant surface sites are
available for adsorption during the initial stage, and
after a lapse of time, it is difcult to occupy the
remaining vacant surface sites due to repulsive
forces between dye molecules adsorbed on the solid
and and those in the solution phase.
The contact time and adsorption rate are
dependent on the initial dye concentration. Gibbs
et al. [81] observed that increasing the initial dye
concentration increased the time required to achieve
complete recovery of the dye. As also expected,
decreasing the adsorbent particle size leads to a
decrease in the time required to reach the equilibrium and a strong increase in the initial adsorption
velocity. The modication of chitosan by grafting

ARTICLE IN PRESS
424

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

reactions [72,73] allows the effect of particle size to


be reduced. It is suspected that these variations
are caused by the specic surface area of the size
fractions, the wide size dispersion around the
median value and the effect of diffusion mechanisms. The equilibrium time increases with the
crosslinking ratio [62,69,71,84,85,95]: GLU crosslinking involves the formation of new interchain
linkages and a loss of chain exibility, and thus
some restriction at the entrance to the polymer
network. Therefore, the extent of crosslinking is
expected to play a great part in adsorption/diffusion
control, especially intraparticle diffusion. However,
the differences are not as marked [62,79]: it is
difcult to nd a homogeneous trend in the
intraparticle diffusion coefcients with increasing
crosslinking ratio. Cestari et al. [61] also indicated
that the adsorption behavior of anionic dyes was
directly related to the dimensions of the dye organic
chains, the amount and position of the sulfonate
groups, and the adsorption temperature. Maghami
and Roberts [50] previously reported that equilibrium was reached more rapidly with the smallest
dye, being attained in less than 2 h with AO 7 while
approximately 9 h were required with AR 27.
However, they indicated that the ionic charge on
the dye appeared to have a negligible effect on the
time to equilibrium.
4.3.4. Effect of stirring rate
Stirring is an important parameter in adsorption
phenomena, inuencing the distribution of the
solute in the bulk solution and the formation of
the external boundary lm. Generally, the rate of
dye removal is inuenced by the degree of agitation
and the uptake increased with stirring rate. The
degree of agitation reduced the boundary-layer
resistance and increased the mobility of the system.
Increase in agitation by increasing stirrer speed
lowers the external mass transfer effect. Uzun and
Guzel [114] reported that the adsorption of the dyes
O II and CV by chitosan must be studied at high
shaking rate. In another recent work [112], they
indicated that there is a small effect of shaking rate
on the adsorption of RY 2 and Rb 5 by chitosan.
Wu et al. [91] also noted that agitation had little
effect on adsorption.
4.3.5. Effect of dryness
To nd the effect of dryness of beads on the
adsorption rate, Chiou and Li [68] used dried beads
to evaluate the adsorption behavior. They reported

that the adsorption rate for wet beads is much faster


than that of dry beads and the time lag to reach
similar adsorption capacity is lower because it takes
time for the dry beads to swell before adsorption
can take place. Chang and Juang [86] also reported
that the adsorption capacity of activated chitosan
with the wet composite beads was generally higher
than that with the dried beads, possibly because of
their stronger afnity of water for the bead matrix.
4.4. Chemistry of the dye
In liquid-phase adsorption, the adsorption capacities of an adsorbent are commonly attributed to
many factors such as its origin, physical, chemical
and mechanical properties, solution conditions, etc.
Dyedye interactions also play an important role as
well as those between dye and aqueous solution. As
described throughout this review, chitosan is being
studied extensively as an adsorbent for dye removal.
However, there is still much to be accomplished in
understanding its mechanisms. In particular, dye
molecules have many different and complicated
structures. This is one of the most important factors
inuencing adsorption and mechanisms. In addition, in most studies, experimental procedures do
not take into account the change in pH due to
chitosan addition and its effect on dye chemistry,
leading to inaccurate interpretation of adsorption
properties. Moreover, changing the experimental
conditions (i.e. the pH, the dye concentration and
the matrix of the solution) can considerably affect
the distribution of dye molecules and consequently
their ability to interact with chitosan. However, to
the best of our knowledge, the comparative effects
of important variables such as the kind of dye, its
pKa, the differences in chemical dye structures on
adsorption behavior on chitosan beads have been
little studied, excepted in recent years.
Guibal et al. [82], studying the adsorption of acid,
direct, mordant and reactive dyes on chitosan,
reported that both adsorption capacities and
kinetics depended on the type of dye involved.
However, any attempt to correlate adsorption
performance to the structure of the dye failed. They
found the following order for anionic dyes: MO
104RB 54AV 54AG 254AB 14AY 254DB
714MB 29 for raw chitosan and AV 54AG
254AY 254MO 104MB 294AB 14DB 71 for
crosslinked chitosan. Following to Maghami and
Roberts [50], who postulated a 1:1 stoichiometry for
the interaction of sulfonic acid groups on the dyes

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

with the protonated amine groups of the chitosan


for mono-, di- and tri-sulfonated dyes, Guibal et al.
[82] attempted to correlate the adsorption capacities
of the dyes to their sulfonate content. They observed
that no meaningful correlation was apparent. Some
dyes containing numerous sulfonic acid groups
exhibited adsorption capacities lower than other
dyes containing only one group per molecule. The
prediction of adsorption performance from only
sulfonic acid content was not possible and other
parameters (pKas, presence of other functional
groups, hydrophobicity) may control adsorption
capacities. They concluded that the differences
between the dyes may be due to different pKas
and/or to the contribution of other interactions in
the adsorption mechanism such as hydrophilic and
hydrophobic interactions due to the different
chemical dye structures. There was no correlation
between the size of the dye and the better efciency
of chitosan for its adsorption compared to CAC. It
is also important to point out that some dyes such
as RB 5 might be subject to hydrolysis, and this
would explain the fact some sulfonic acid groups are
not able to react with chitosan [80].
Different conclusions have been reported by
Wong et al. [99] who attributed the differences in
the degree of adsorption mainly to the chemical
structure of each dye. A detailed study of the order
of afnity of ve acid dyestuffs for ion-exchange
onto chitosan was reported by the authors. They
noted that there was a great variation in the afnity
of chitosan for these dyes in a single solute system.
The order of extent of decolorization was AO
124AO 104AR 734AR 184AG 25. The differences in adsorption capacities may be due to the
effect of molecular size and the number of sulfonate
groups of each dye. They concluded that monovalent and/or smaller dye molecules have greater
adsorption capacities due to an increase in the dye/
chitosan ratio in the batch system. The smaller dye
molecules are able to undertake a deeper penetration of dye into the internal pore structure of the
chitosan particles. A similar interpretation was
previously given by McKay et al. [44] for the
adsorption of basic dyes on chitin. Smith et al. [52]
reported that the molecular size of the dye was a
major factor in adsorption characteristics. They
noted that small, low MW dyes adsorbed best on
chitosan. Crini et al. [72,73] also found signicant
variations in afnity of the different cationic dyes
(BB3 and BB 9) to grafted chitosan, with minimum
adsorption for BB 3 (166.5 mg/g) and maximum for

425

BB 9 (121.9 mg/g). It was found that the molecular


size of the dye was a major factor in adsorption
characteristics. Juang et al. [93] observed that the
adsorption capacity, dened as the amount at
the isotherm plateau, depended on the nature of
the dye molecules. Under comparable experimental
conditions, the capacity decreased in the order RR
2224RY 1454RB 222, conrming the role of dye
structure. They also added that the adsorption rate
also depended on the dye chemistry and followed
the same order. Cestari et al. [61] also indicated that
the adsorption behavior of dyes was inuenced by
the chemical structure of the dye molecules, but they
reported that the link with the dye structure was not
clearly identied.
4.5. Effect of the solution conditions
4.5.1. Effect of pH
The pH of the dye solution plays an important
role in the whole adsorption process and particularly on the adsorption capacity, inuencing not
only the surface charge of the adsorbent, the degree
of ionization of the material present in the solution
and the dissociation of functional groups on
the active sites of the adsorbent, but also the
solution dye chemistry. It is important to indicate
that while the adsorption on CAC was largely
independent of the pH, the adsorption of (anionic)
dyes on chitosan was controlled by the acidity of the
solution.
pH affects the surface charge of the adsorbent.
Chitosan is a weak base and is insoluble in water
and organic solvents, however, it is soluble in dilute
aqueous acidic solution (pHo6.5) which can convert the glucosamine units into a soluble form
R-NH+
3 . It gets precipitated in alkaline solution or
with polyanions and forms gels at lower pH. Its pKa
depends on the DD, the ionic strength and the
charge neutralization of amine groups. In practice it
lies within 6.56.7 for fully neutralized amine
functions [32]. So, chitosan is polycationic in acidic
medium: the free amino groups are protonated and
the polymer becomes fully soluble and this facilitates electrostatic interaction between chitosan and
the negatively charged anionic dyes. This cationic
property will inuence the adsorption procedure,
especially in the case of anionic dyes, depending on
the charge and functions of the dye under the
corresponding experimental conditions [82]. In
1994, Muzzarelli and co-workers [53,54], studying
the interactions of dyes with chitosan in solution by

ARTICLE IN PRESS
426

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

circular dichroism analysis, observed that the pH of


the solution changed both the extent and the mode
of the binding. They explained this change in the
mechanism by changes in the conformation of
the polymer chains.
In the literature, the ability of the anionic dyes to
adsorb onto chitosan beads is often attributed to the
surface charge which depends on the pH of the
operating batch system, as mentioned. Dye adsorption occurred through electrostactic attraction on
protonated amine groups and numerous workers
concluded that the inuence of the pH conrmed
the essential role of electrostatic interactions between the chitosan and the target dye. For example,
Chatterjee et al. [83] indicated that chitosan had a
positively charged surface below pH 6.5 (point of
zero potential), and reducing the pH increased the
positivity of the surface, thus making the adsorption
process pH sensitive. Decreasing the pH makes
more protons available to protonate the amine
group of chitosan with the formation of a large
number of cationic amines. This results in increasing
dye adsorption by chitosan due to increased
electrostatic interactions. Differences in pH of the
solution have also been reported by Gibbs et al.
[80,81] to inuence the dye adsorption capacity of
chitosan and its mechanism. They noted that, at low
pH, chitosans free amino groups are protonated,
causing them to attract anionic dyes, demonstrating
that pH is one of the most important parameters
controlling the adsorption process. Crini et al.
[72,73] also found that the adsorption capacity
of cationic dyes on chitosan-grafted materials
was strongly affected by the pH of solution and
was generally signicantly decreased by increasing
the pH.
pH is also known to affect the structural stability
of dye molecules (in particular the dissociation of
their ionizable sites), and therefore their color
intensity. For example, BG 4, a cationic dye
(pKa 10.3) gets protonated in acidic medium
and deprotonated in basic medium. Consequently,
the dye molecule has high positive charge density at
a low pH. This indicates that the deprotonation
(or protonation) of a dye must be take into account.
If the dyes to be removed are either weakly acidic
or weakly basic, then the pH of the medium affects
their structure and adsorption. Initial pH also
inuences the solution chemistry of the dyes:
hydrolysis, complexation by organic and/or inorganic ligands, redox reactions, and precipitation are
strongly inuenced by pH, and on the other side

strongly inuence speciation and the adsorption


availability of the dyes.
Two interesting experimental facts must be
pointed out. Firstly, the free amine groups in
chitosan are much more reactive and effective for
chelating pollutants than the acetyl groups in chitin,
and there is no doubt that amine sites are the main
reactive groups for (anionic) dye adsorption, though
hydroxyl groups (especially in the C-3 position) may
contribute to adsorption. Almost all functional
properties of chitosan depend on the chain length,
charge density and charge distribution and much of
its potential as biosorbent from its cationic nature
and solution behavior. However, at neutral pH,
about 50% of total amine groups remain protonated and theoretically available for the adsorption
of dyes. The existence of free amine groups may
cause direct complexation of dyes co-existing with
anionic species, depending on the charge of the dye.
As the pH decreases, the protonation of amine
groups increases together with the efciency.
The optimum pH is frequently reported in the
literature to be around pH 36 (Table 10). Below
this range, usually a large excess of competitor
anions limits adsorption efciency. This competitor
effect is the subject of many studies aiming to
develop materials that are less sensitive to the
presence of competitor anions and to the pH of the
solution, as described in the next two paragraphs.
Secondly, it is not really the total number of free
amine groups that must be taken into account but
the number of accessible free amine groups. There
are several explanations for this. The availability of
amine groups is controlled by two important
parameters: the crystallinity of polymer and the
diffusion properties. It is known that some of
the amine sites on chitosan are involved in both
the crystalline area and in inter or intramolecular
hydrogen bonds. Moreover the residual crystallinity
of the polymer may control the accessibility to
adsorption sites. The DD also controls the fraction
of free amine groups available for interactions with
dyes. Indeed, the total number of free amine groups
is not necessarily accessible for dye uptake. Actually, rather than the fraction or number of free
amine groups available for dye uptake, it would be
better to consider the number of accessible free
amine groups. Guibal et al. [79] recently showed
that not all the amino groups are really available or
at least accessible. They also concluded that the
hydrogen bonds linked between monomer units of
the same chain (intramolecular bonds) and/or

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

between monomer units of different chains (intermolecular bonds) decrease their reactivity. The
weakly porous structure of the polymer and its
residual crystallinity are critical parameters for the
hydratation and the accessibility to adsorption sites.
4.5.2. Effect of pH variation
To date it remains difcult to establish clear
trends in the adsorption properties of chitosan
materials for the recovery of dyes. The variability in
published results may deter potential users. There
are many reasons for this difculty in comparing
adsorption performances such as differences in
chitosan and above all experimental conditions
(equilibrium time, adsorbent dose), and also underestimation of the inuence of pH, especially pH
variation effects during the process [81]. Indeed,
frequently, batch systems used in the literature do
not take into account the change in pH.
Few studies have been published on the interpretation of the pH variation during adsorption
process. Sakkayawong et al. [109] reported that the
system pH changed during RR 141 adsorption by
chitosan. The explanation for this was that under
acidic conditions hydrogen atoms in the solution
could protonate the amine groups of the polymer
and thus causes the increased pH. In addition, the
adsorption efciency was systematically greater for
solutions whose pH was controlled during the
adsorption than for solutions for which pH varied
along the uptake process. At increasing adsorbent
dosage, Gibbs et al. [81] observed that the addition
of chitosan to the solution strongly increased its pH
and noted that the pH variation increased exponentially. The authors interpreted this phenomenon as
related to the increase of the number of amine
groups available for protonation, consequently
causing a marked increase in the pH. Increasing
adsorbent dose also reduced polymer saturation.
To diminish the inuence of pH change on the
interpretation and modeling of adsorption performance, chitosan can be conditioned by contact with
a solution of sulfuric acid [81] or other chemicals
[83], as shown below.
4.5.3. pH sensitivity
It is well known that chitosan hydrobeads lose
their integrity as a result of partial dissolution in
acidic solvent, since protonation of amine groups
causes the polymer to dissolve, making them
unsuitable for reuse. Crosslinking of the chitosan
beads has been proposed but can result in poor

427

adsorption. Thus, it is of interest to increase the


integrity of the beads as well as adsorption properties at acidic pH, and also to diminish the inuence
of the pH change during adsorption process.
Recently, several authors [72,81,83] chose to
modify raw chitosan without crosslinking in order
to diminish the pH variation. Crini et al. [72]
proposed the use of new chitosan derivatives as
adsorbents in acidic medium. The materials are
prepared by grafting reactions followed by a
solubilization/reprecipitation step, converting the
chitosan akes into a powder having an advantageous specic surface area. The originality of the
materials is their inversed pH domain of solubility
compared to the parent polymer. This characteristic
allowed adsorption capacity to be studied in strong
aqueous acidic media (pH 2/3), without further
crosslinking or pH sensitivity. Gibbs et al. [81]
showed that the preliminary protonation of amine
groups, obtained by contact with a sulfuric acid
solution, reduces the variation of solution pH
following adsorbent addition. In this case, pH
variation during the adsorption process was signicantly lower. A direct correlation can be
observed between the theoretical neutralization
curve of chitosan, calculated from the pKa, and
the adsorption performance at different pHs. They
also observed that when the initial dye concentration increased, pH variation increased, indicating
that the protonation of the dye plays an important
role in this variation. However, as previously
mentioned, this pretreatment (preprotonation)
strongly reduced adsorption performance at both
equilibrium and kinetic levels. In particular, the
time necessary to reach equilibrium increased up to
threefold depending on the experimental conditions.
Chatterjee et al. [83] proposed the conditioning of
the beads with ammonium sulfate to reduce the pH
sensitivity of the process with interesting adsorption
properties in the case of eosin Y. This modication
would probably form complexes between cationic
amino groups and ammonium sulfate, which undergoes interactions with dye anions.
4.5.4. Effect of ionic strength
For almost all the treatment strategies, an
important factor which has not yet been adequately
characterized is the effect of typical wastewater
contaminants on decolorization efciencies [21]. In
typical dyeing systems it is well known that certain
additives such as salts and surfactants can either
accelerate or retard dye adsorption processes. For

ARTICLE IN PRESS
428

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

example, sodium chloride is often used as a


stimulator in dyeing processes. Salts may have two
functions: (i) they may screen the electrostatic
interaction of opposite charges in adsorbents and
the dye molecules, and an increase in salt concentration could decrease the amount of dye adsorbed;
(ii) they may enhance the degree of dissociation
of the dye molecules and facilitate the amount of
pollutant adsorbed. The former function seems to
be dominant in the literature, conrming that
electrostatic interaction is responsible for adsorption of acid dyes by chitosan. The situation becomes
much more signicant when both the adsorbate and
adsorbent are charged bodies. So, the ionic strength
may be another important factor in the adsorption
of certain dyes onto chitosan [68,72,95,100,105].
Crini et al. [72] discussed the strong effect of
sodium chloride on the adsorption process. They
concluded that the capacities depended on the ionic
strength of the solution. Added salts affected
adsorption via two mechanisms, either by screening
the coulombic potential between the adsorbing molecule and charged adsorbents, or by adsorbing
preferentially on the active sites of the adsorbent.
They also noted that the increase in NaCl concentration reduced the dose of chitosan necessary to
adsorb the pollutant. In another work, Martel et al.
[95] reported that the addition of salts in batch
systems diminished the solubility of the dyestuff in
the solution and thus favored its precipitation onto
polymer, suggesting the presence of an aggregation
mechanism increasing the adsorption capacity in the
presence of salts. Miyata et al. [100] also conrmed
the fact that the addition of sodium chloride greatly
affected the adsorption of acid and direct dyes by
chitosan in agreement with the prediction of the
chemisorption mechanism. Shimizu et al. [105]
noted that the addition of sodium chloride greatly
affected the adsorption of AR 1 by crosslinked
chitosan while no effect was observed on the
adsorption of AR 138, indicating the role of the
dye structure. They concluded that the adsorption
of AR 138 also involved hydrophobic interactions
between the alkyl groups in the dye and the
hydrophobic functions in the adsorbent. However,
these interactions are strongly inuenced by the pH
of the solution. Chiou and Li [68] also reported that
addition of inorganic salts to the adsorption system
is an effective way of inuencing adsorption.
However, they observed that the presence of NaCl
in the batch solution lowers the adsorption capacity,
slows down the initial adsorption rate and increases

the equilibrium time. They suggested that the


addition of NaCl reduced the electrostatic interaction between dye and crosslinked chitosan.
4.5.5. Effect of competitive molecules and ions
Competitive adsorption occurs where the adsorption of a mixture of adsorbates is carried out on one
surface. Some of the components in the efuent may
induce the adsorption of others or may coadsorb
along with another components. As mentioned
above, the variability of wastewater must be taken
into account in the design of any decolorization
system. However, only very limited information is
available on the competitive adsorption of dye
molecules with chitosan-based materials.
In an effort to further understand the adsorption
process of dye molecules onto chitosan, Chiou et al.
[68,69] evaluated the effect of mixed-dye solutions.
They reported that the presence of other molecules
might affect the adsorption of a particular molecule.
They concluded that several dyes in a solution
would compete against each other for available
sites. Those having the greatest ionic potential
would be removed rst, and if the sites were still
undersaturated, then those having lower ionic
potential would be removed in sequence. The more
electronegative molecules are attracted to the surface more strongly. Although the presence of more
than one dye in a solution creates competition for
adsorption sites, the total adsorption capacity has
been found to increase. The presence of counterions, interferents, and/or other substances in solution affects the adsorption of dyes by chitosan, as
recently reported by Wen et al. [116]. They noted
that the presence of appreciable quantities of Na+
did not have any effect on RR 195 removal by
chitosan. Addition of other cation species (Ca+,
Mg2+, Fe2+) decreased color removal. This was
attributed to chelation between cations and chitosan
chains, which decreased the electrostatic interaction
between RR 195 and chitosan. Compared with
Fe2+ alone, the combination of Fe2+/HCO
3
increased color removal. However, the mechanisms
of combination need to be explored.
4.6. Stability
In view of industrial developments of the various
kinds of chitosan derivatives, the stability of the
materials is of utmost importance. Unfortunately,
in this area, the literature produces very little
information except in recent publications [18,62,72].

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

Being a biopolymer, chitosan is biodegradable.


This property may be a serious drawback for longterm applications in adsorption processes, in particular in dynamic systems using xed-bed columns.
The low thermal stability of chitosan and its
degradation resulting from acidic hydrolysis are
another important criterion to take into account.
Generally, it is possible to reinforce the chemical
stability of the chitosan by crosslinking treatments
or grafting reactions [62,68,73,84,88]. However,
despite the number of studies dealing with chitosan
gel beads, their mechanical stability is signicantly
less documented [18]. Guibal [18] pointed out the
role of crosslinking reactions on chitosan bead
stability. He reported that, in the case of GLU
treatment, the beads obtained lose their elasticity
and deformability, and under strong strain the
materials break and small granules are formed. He
concluded that this was a serious drawback for
large-scale applications.
4.7. Desorption of dyes
Polymeric adsorbents present considerable advantages such as their high adsorption capacity,
selectivity and also the facility of regeneration. The
regeneration of the adsorbent may be crucially
important for keeping the process costs down and to
open the possibility of recovering the pollutant
extracted from the solution. For this purpose, it is
desirable to desorb the adsorbed dyes and to
regenerate the chitosan derivative for another cycle
of application. Generally, the regeneration of
saturated chitosan for non-covalent adsorption
can be easily achieved by using an acid solution as
the desorbing agent.
Chatterjee et al. [83] proposed to desorb the dye
from the beads by changing the pH of the solution
and they showed that the beads could be reused ve
times without any loss of mechanical or chemical
efcacy. The rate of desorption was found to
increase with an increase in pH of the eluent. They
concluded that the chitosan was recyclable. This is
interesting because adsorption processes could be
considered as potential methods for the decontamination of the efuents of textile industries since dyes
can be selectively adsorbed, concentrated, and then
recycled. Though CAC has received a great deal of
attention, this material is frequently non-selective
and difcult to reuse [82]. Hu et al. [110] reported
that the dyes adsorbed on chitosan nanoparticles
could be desorbed in an alkaline medium (causing

429

the de-aggregation of the particles) but not in a


neutral medium. Trung et al. [108] showed that
decrystallized chitosan can be regenerated by
sufuric acid and was reusable more than 10 times.
They hypothesized that the protons of the sulfuric
acid are effective in reducing the dissociation of the
anionic groups of the dyes. Lima et al. [58] reported
that grafted chitosan can be regenerated with
aqueous solutions containing sodium chloride without the use of organic solvents or pH modication.
Since the interaction between BB 9 and grafted
chitosan are driven mainly by chemisorption,
organic solvents could be good candidates for the
regeneration of the materials, as suggested by Crini
et al. [72]. After saturation, the materials are easily
regenerated using ethanol as extraction solvent. The
authors indicated that the adsorption capacities
remained unchanged after this treatment: this
showed both the chemical stability of the materials
and reproducibility of the values. Guibal [18]
pointed out that adsorbent recycling was necessary
for making the use of chitosan cost-efcient for
environmental applications. However, as yet there is
little literature on this topic.
5. Adsorption mechanisms
The rst major challenge for the adsorption eld
is to select the most promising types of adsorbent,
mainly in terms of efciency and low cost. The next
real challenge is to clearly identify the adsorption
mechanism(s), in particular the interactions occurring at the adsorbent/adsorbate interface.
Two mechanisms are clearly established for the
interpretation of metal adsorption on chitosan
materials, i.e. electrostatic interactions in acid media
(ion-exchange) and metal chelation (coordination),
although the formation of ion pairs has also been
reported [18,19,21,22]. Metal ion adsorption is
assumed to occur through single or mixed mechanisms including coordination on amino groups in a
pendant fashion or in combination with vicinal
hydroxyl groups, and ion-exchange with protonated
amino groups through proton exchange or anion
exchange, the counter ion being exchanged with the
metal anion. The nature of the reaction depends
upon several parameters related to the adsorbent
(ionic charge), to the solution (pH) and the
chemistry of the metal ion (ionic charge, ability to
be hydrolyzed and to form polynuclear species). For
more details on these mechanisms, two recent
reviews can be consulted [18,19].

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

430

limitations due to heat transfer can be neglected.


Providing sufcient stirring to avoid particle and
solute gradients in the batch system also makes it
possible to ignore bulk diffusion, which can be
considered instantaneous. So, the most important
steps are lm diffusion, pore diffusion and chemical
reaction.
Previously, several studies [4956] showed that
amine sites were the main reactive groups for dyes,
though hydroxyl groups (especially in the C-3
position) might contribute to adsorption, and the
intermolecular interactions of the dye molecules are
most probable in the chitosan-dye systems. The
occurrence of an interaction in aqueous solutions
between the hydrophilic chitosan and the anionic
hydrophobic dyes was proved by Muzzarelli and coworkers [53,54] using optical and thermodynamic
techniques. It is now recognized that chemisorption
(ion-exchange, electrostatic attractions) is the most
prevalent mechanism with the pH as the main factor
affecting adsorption. Chemisorption, a strong type
of adsorption in which molecules are not exchanged
but electrons may be exchanged, is commonly cited
as the main mechanism for the adsorption of
anionic dyes in acidic conditions. Scheme 1 briey
describes the mechanism: in the presence of H+, the
amino groups of chitosan become protonated; also,
in aqueous solution, the anionic dye is rst dissolved
and the sulfonate groups in the case of acid or
reactive dyes dissociate and are converted to anionic
dye ions; the adsorption process then proceeds due
to the electrostatic attraction between these two
counterions. In general, as the initial dye concentration increases, the equilibrium pH decreases. This is
consistent with the principles of ion-exchange since
as more dye molecules are adsorbed onto material,
more hydrogen ions are released, thereby decreasing
the pH.

On the contrary, for dye molecules, the mechanisms


by which adsorption onto chitosan occurs has been a
matter of considerable debate with surface adsorption,
chemisorption, diffusion and adsorption-complexation
being the prevalent theories. Among the large number
of papers dedicated to the removal of dyes by chitosanbased materials, most focus on the evaluation of
adsorption performances and only a few aim at
gaining a better understanding of adsorption mechanisms. Different studies have reached different conclusions. This can perhaps be explained by the fact that
different kinds of interactions such as chemical
bonding, ion-exchange, hydrogen bonds, hydrophobic
attractions, van der Waals force, physical adsorption,
aggregation mechanisms, dyedye interactions, etc.,
can act simultaneously. It is important to note that
variation in chitosan preparation and actual methodology often adds to the complication. Wide ranges of
chemical structures, pH, salt concentrations and the
presence of ligands also makes the comparison of
results difcult.
In general, the mechanism for dye removal by
adsorption on an adsorbent material may be
assumed to involve the following four steps:
(i) bulk diffusion: migration of dye from the bulk
of the solution to the surface of the adsorbent;
(ii) film diffusion: diffusion of dye through the
boundary layer to the surface of the adsorbent;
(iii) pore diffusion or intraparticle diffusion: transport of the dye from the surface to within the
pores of the particle;
(iv) chemical reaction: adsorption of dye at an active
site on the surface of material via ion-exchange,
complexation and/or chelation.
Heat transfer may, in theory, be added. However,
due to the heat transfer properties of water, kinetic
H+

NH3+

NH2
protonation

chitosan

H2O
dye

SO3Na

dye

SO3-

Na+

dissociation
NH3+
chitosan

+ dye

SO3-

NH3+ -O3S

dye

electrostatic interaction

Scheme 1. Mechanism of anionic dye adsorption by chitosan under acidic conditions.

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

The Scheme 1 is accepted by numerous authors as


the main adsorption mechanism. Juang and coworkers [8693] carried out different studies to
challenge the theory that amino groups were
primarily responsible for dye binding in crosslinked
chitosan. Their investigations centered on the
adsorption of anionic dyes by chitosan. They
prepared several types of chitosan including akes,
beads and composite materials, and found a strong
correlation between the functional groups present
on the adsorbent surface and their adsorption
capacity for dye molecules. Since chitosan was
soluble in water at acid pH, the electrostatic
properties of chitosan were pH responsive and the
pH played an important role in chitosan-based
adsorbent processes. They concluded that the
materials had particularly high selectivity with
respect to the extraction of dyes by ion-exchange.
Sakkayawong et al. [109] reported that the mechanism of adsorption of RR 141, a reactive dye, by
chitosan under acidic conditions was by chemisorption, while under caustic conditions it was both by
physical interaction (rapid surface adsorption) and
chemical adsorption. Surface adsorption is another
mechanism by which dye molecules may be bound
to chitosan. This mechanism is a surface reaction
where a molecule is attracted to a charged surface
without the exchange of ions or electrons. Other
studies have found evidence that chitosan takes up
dyes by surface adsorption. Uzun and Guzel [115]
reported that both physical adsorption and chemisorption occurred simultaneously between O II and
chitosan. They concluded that physical surface
interaction occurring with multilayer adsorption,
played an important role in the adsorption mechanism. In another work [112], they reported that,
because the BET specic surface area of chitosan
was very low (0.65 m2/g), mechanisms were mainly
controlled by surface diffusion, and at lower
temperatures, surface diffusion was even more
dominant. Using kinetic studies, they showed that
the mechanism of action was surface adsorption
rather than ion-exchange. On the contrary, Cestari
et al. [62] arrived at different conclusions. They
pointed out the importance of site accessibility and
that the low specic surface areas implied that
purely physical adsorption onto the surface of the
beads was not signicant.
Other authors [72,73,80,81,83,91,95,102,103] concluded that the uptake of dyes on ake- and beadtypes of chitosan, and grafted chitosan may proceed
mainly through ion-exchange mechanisms. Gibbs

431

et al. [81] reported that the adsorption of AG 25 on


raw and preprotonated chitosan consisted of two
main steps: a rapid surface adsorption followed by
diffusion and chemisorption of dye molecules in the
polymer network. In another work, they also
indicated that RB 5 was bound to chitosan not
only through electrostatic attractions but also a dye
aggregation mechanism which can play a role under
certain experimental conditions [80]. Crini et al.
[72,73], studying the adsorption of cationic dyes by
grafted chitosan, conrmed that the mechanism was
chemisorption of the dye via the formation of
electrostatic interactions. The adsorption phenomenon mainly depends on the interaction between the
surface of the grafted chitosan and the adsorbed
species. However, they added that the mechanism
was also due to physical surface adsorption and
hydrogen bonding because of the polymer network.
Adsorption increases as the surface area of the
adsorbent increases. They concluded that the
mechanism was a multistep complex process since
other interactions such as diffusion and hydrophobic and steric interactions could play an important
role. Martel et al. [95] studied the adsorption of
acid, direct, mordant and reactive dyes on raw
chitosan, chitosan beads and cyclodextrin-grafted
chitosan. They observed that the materials did not
interfere with the dye molecule in the same manner
and suggested the presence of different interactions
in the adsorption mechanism simultaneously. Saha
et al. [102], studying the adsorption of an azo dye
onto chitosan akes by kinetic studies concluded
that the adsorption mechanisms were both transport- and attachment-limited. In other words,
the mechanism is separated in two processes: the
transport of dye from the bulk solution to the
surface of the akes followed by the attachment of
the dye to chitosan by chemisorption. Wu et al. [91]
also showed that intraparticle diffusion plays an
important role in the adsorption mechanism,
together with ion-exchange mechanisms. Similar
conclusions have been reached in the case of
crosslinked chitosan beads. Chiou et al. [67]
demonstrated that the strong electrostatic interactions in acidic medium between the NH+
3 groups
present in crosslinked materials and dye anions can
explain the adsorption mechanism. Hebeish et al.
[84,85] also indicated that crosslinked materials
possesses greater afnity for the adsorption of
reactive and direct dyes than for basic dyes. They
indicated that the anionic character of the former
dyes was responsible for the greater dye adsorption,

ARTICLE IN PRESS
432

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

unlike the basic dye with its cationic character,


conrming the chemisorption mechanism.
Evidence has also been found that aggregation, a
strong type of interaction depending on the pH, can
be involved in chitosan-dye binding. The aggregation mechanism affects the size of the dye particles
and their ability to diffuse into the internal porous
network. Hu et al. [110] reported that nanoparticles
quickly aggregated after interacting with the dye
molecules, suggesting the replacement of the hydrogen bonds between polymer chains by electrostatic
interactions between dyes and chains. So, an
aggregation mechanism could be included. Increasing the dye concentration inuences the aggregation
mechanism as suggested by Gibbs et al. [80]. The
size of the dye aggregate can inuence its diffusivity,
especially for low-porosity chitosans. The aggregation phenomenon is enhanced by the presence of an
electrolyte in the solution and by ionic strength.
Hydrophilic [53,54] and hydrophobic [51,72,73,105]
interactions, depending on the chitosan structure,
have been also proposed by several authors. For
example, Seo et al. [51] previously pointed out the
predominant contribution of hydrophobic interactions. Shimizu et al. [105] also showed that the
adsorption of AR 138 by crosslinked chitosan
involved hydrophobic interactions between the
(hydrophobic) dye and the hydrophobic functions
in the polymer network. Crini et al. [72,73] reported
that dye adsorption on grafted chitosan in solution
was essentially an exchange process and molecules
adsorbed not only because the were attracted by
materials but also because the solution might reject
them due to the presence of strong dyedye
hydrophobic interactions.
The difference in the degree of adsorption may
also be attributed to the chemical structure of each
dye, as proposed by Wong et al. [9799]. They
demonstrated that the dye molecules, when adsorbed on chitosan, are more or less attached to
chitosan chains in a at or layered manner, that is,
covering long chitosan macromolecules with benzene rings oriented parallelly (as far as possible) to
the polyamine chain of chitosan. If the attachment
of the dye were at one point only (electrostatic
reaction between amino and sulfonate groups), the
dye molecule would be expected to be more spatially
oriented. This result conrms that, in addition to
electrostatic binding, there is a strong possibility of
hydrogen bonding between chitosan and dye
molecules [99]. Gibbs et al. [80] also reported that
dye molecules with phenyl groups can adopt a

planar structure and readily tend to form intermolecular interactions that facilitate permanent
aggregation under certain experimental conditions
(presence of surfactant, pH).
Numerous authors concluded that the binding
was a chemisorption reaction and the adsorption
phenomenon mainly depended on the interactions
between the surface of the adsorbent and the
adsorbed species. However, all the studies arrive at
contrasting conclusions showing the difculty of
using simple models for the interpretation of the
interactions of these polymeric materials with dyes.
Much work is necessary to clearly demonstrate the
adsorption mechanism for the different types of
chitosan-based materials.
6. Modeling
6.1. Equilibrium isotherm models
Adsorption properties and equilibrium data,
commonly known as adsorption isotherms, describe
how pollutants interact with adsorbent materials
and so, are critical in optimizing the use of
adsorbents. In order to optimize the design of an
adsorption system to remove dye from solutions, it
is important to establish the most appropriate
correlation for the equilibrium curve. An accurate
mathematical description of equilibrium adsorption
capacity is indispensable for reliable prediction of
adsorption parameters and quantitative comparison
of adsorption behavior for different adsorbent
systems (or for varied experimental conditions)
within any given system.
Adsorption equilibrium is established when the
amount of dye being adsorbed onto the adsorbent is
equal to the amount being desorbed. It is possible to
depict the equilibrium adsorption isotherms by
plotting the concentration of the dye in the solid
phase versus that in the liquid phase. The distribution of dye molecule between the liquid phase and
the biosorbent is a measure of the position of
equilibrium in the adsorption process and can
generally be expressed by one or more of a series
of isotherm models [122126]. The shape of an
isotherm may be considered with a view to
predicting if a sorption system is favorable or
unfavorable. The isotherm shape can also provide qualitative information on the nature of the
solutesurface interaction. In addition, adsorption
isotherms have been developed to evaluate the
capacity of chitosan materials for the adsorption

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

of a particular dye molecule. They constitute the


rst experimental information, which is generally
used as a convenient tool to discriminate among
different materials and thereby choose the most
appropriate one for a particular application in given
conditions.
The most popular classication of adsorption
isotherms of solutes from aqueous solutions has
been proposed by Giles et al. [125,126]. Four
characteristic classes are identied, based on the
conguration of the initial part of the isotherm
(i.e., class S, L, H, C). The subgroups relate to the
behavior at higher concentrations. The Langmuir
class (L) is the most widespread in the case of
adsorption of dye compounds from water, and it is
characterized by an initial region, which is concave
to the concentration axis. Type L also suggests that
no strong competition exists between the adsorbate
and the solvent to occupy the adsorption sites.
However, the H class (high afnity) results from
extremely strong adsorption at very low concentrations giving rise to an apparent intercept on the
ordinate. The H-type isotherms suggest the uptake
of pollutants by materials thgrough chemical forces
rather than physical attraction.
There are several isotherm models available for
analyzing experimental data and for describing the
equilibrium of adsorption, including Langmuir,
Freundlich, BET, Toth, Temkin, Redlich-Peterson,
Sips, Frumkin, Harkins-Jura, Halsey, Henderson

433

and Dubinin-Radushkevich isotherms. These equilibrium isotherm equations are used to describe
experimental adsorption data. The different equation parameters and the underlying thermodynamic
assumptions of these models often provide insight
into both the adsorption mechanism, and the
surface properties and afnity of the adsorbent.
Therefore, it is important to establish the most
appropriate correlation of equilibrium curves to
optimize the condition for designing adsorption
systems.
Various researchers have used these isotherms to
examine the importance of different factors on dye
molecule sorption by chitosan. However, the two
most frequently used equations applied in solid/
liquid systems for describing sorption isotherms are
the Langmuir [123] and the Freundlich [124] models
and the most popular isotherm theory is the
Langmuir one which is commonly used for the
sorption of dyes onto chitosan (see Table 8). Table 9
reports the corresponding equations that can be
used for tting experimental data. The symbols used
in the equations are dened in the Nomenclature
section. Linear regression was frequently used to
determine the most frequently used model throughout the years.
The Langmuir model was found to be the most
appropriate to describe the adsorption process in
the case of (i) RB 5 [79,81], BB 9 [73], DB [75], eosin
[83], AG 27 [110], O II [115] and CV [115] on

Table 9
The two most popular equilibrium isotherm equations and their linear forms and equation parameters
Isotherm

Equation

Langmuir

qe

x
K LCe

m 1 aL C e

Assumptions

 Monolayer adsorption
 The sorption takes place at specic sites within the

Linear form
Ce
1
aL

Ce
KL KL
qe

adsorbent

 Once a dye molecule occupies a site


 The adsorbent has a nite capacity for the adsorbate (at


Freundlich

F
qe K F C 1=n
e

equilibrium, a saturation point is reached where No further


adsorption can occur)
All sites are identical and energetically equivalent
The adsorbent is structurally homogeneous

 Multilayer adsorption
 The model applies to adsorption on heterogeneous surfaces
with interaction between adsorbed molecules

 The adsorption energy exponentially decreases on


completion of the sorptional centres of an adsorbent

 This is an empirical equation employed to describe


heterogeneous systems.

ln qe ln K F

1
ln C e
nF

ARTICLE IN PRESS
434

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

chitosan, (ii) RR 189 [68], RR 222 [69], RO 16 [74]


and AR 138 [105] on crosslinked chitosan, (iii) BB 9
[58], BV 3 [106] and O II [113] on grafted chitosan.
Gibbs et al. [81], studying the adsorption of RB 5 on
chitosan reported that the adsorption isotherms
were characterized by a steep increase in the
adsorption capacity (indicating a great afnity of
the adsorbent for the dye), followed by a plateau
representing the maximum capacity at saturation of
the monolayer. This shape corresponds to the
typical Langmuir-type equation. They concluded
that the adsorption was very favorable and almost
irreversible. Similar conclusions have been reported
by Saha et al. [102] who studied the adsorption of an
azo dye onto chitosan akes. They concluded that
the isotherm was dominated by a monolayer
adsorption process, in which the akes were
protonated and rapidly interacted with the dye to
form complexes in accordance with Scheme 1. They
assumed that the adsorbed layer was one molecule
thick and the sites were homogeneous, conrming
the applicability of the Langmuir model [123].
Sakkayawong et al. [109] demonstrated that RR
141 adsorption on chitosan obeyed the Langmuir
model which suggested that biosorption occurred
on the homogeneous surface. Wong et al. [98,99]
reported that the Langmuir model was found to
provide the best theoretical correlation of the
experimental data for the adsorption of ve acid
dyes. They noted the high degree of correlation for
the linearized Langmuir relationship suggested a
single surface reaction with constant activation
energy and was the predominant adsorption step.
However, they also reported that, if the whole
concentration range was divided into three different
regions, excellent ts to the experimental data could
be observed with the Freundlich model, especially at
the lower concentrations [99]. Hu et al. [110],
studying the adsorption of AG 27 on nanochitosan
found that the sorption tted the Langmuir model
well, especially when the concentration was high.
Crini et al. [72,73], studying the adsorption of BB 9
and BB 3 by grafted chitosan, used both the
Langmuir and the Freundlich models to describe
the adsorption equilibrium. It was found that
the data tted better to Langmuir model and error
analysis investigations highlighted the non-linear
method as a better way to obtain the isotherm
parameters. Dos Anjos et al. [63] observed that both
Freundlich and Langmuir models tted IC adsorption on chitosan well, which indicated adsorption
by combined mechanisms onto a heterogeneous

surface. In other cases, the Freundlich equation was


preferred. Kim and Cho [71], studying the adsorption of RB 5 on crosslinked chitosan, reported that
the Freundlich isotherm best tted the data over the
entire pH and temperature range of the solution.
Chang and Juang [86], studying the adsorption of
RR 222, AO 51 and BB 9 onto chitosan, reported
that in a simple way the two-parameter Langmuir
and Freundlich equations can treat the isotherms.
The Freundlich equation was preferred for the
description of the isotherms of reactive dyes and the
isotherms of acidic and basic dyes were better tted
by the Langmuir model, depending on the dye
concentration. However, they conrmed that the
Freundlich was an empirical approach applicable to
the adsorption of single solutes within a xed range
of concentration [124]. This model is generally
suitable for high- and middle-concentration environments and is not suitable for low concentrations
because it does not meet the requirements of
Henrys law [86].
Two important points must be pointed out. The
rst is that, despite its highly idealistic simplicity,
the two-parameter isotherm model remains a useful
and convenient tool for the comparing results from
different sources on a quantitative basis. Many
liquid adsorption studies on chitosan have been
carried out by tting the Langmuir and Freundlich
isotherm parameters to the experimental data as
shown in Table 8. However, while this approach is
convenient for the characterization of data, it has
limited potential only for predicting behavior under
conditions within the ranges of experimental measurements, because the assumptions of these models
are not closely based on the actual adsorption
processes. Both models suffer from the drawback
that equilibrium data over a wide concentration
range cannot be tted with a single set of constants.
In addition, the Langmuir and Freundlich models,
initially developed for modeling gas solutes on
metallic surfaces, are based on the hypothesis of
physical adsorption. In the case of dye adsorption,
which is more chemical than physical, it would be
more appropriate to consider dye adsorption with
models based on chemical reactions, in order to take
into account the real phenomena between the
chitosan material and the dye, and also the dye
dye interactions. Therefore, the three-parameter
isotherm equations such as the Redlich-Peterson
and the Sips models which combine the features of
both Langmuir and Freundlich models, are preferred. These models are still being used and much

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

more work is required. It was demonstrated that


the three-parameter models tted the experimental
data better than the two-parameter models because
they take into account additional parameters
(pH, temperature) and other interactions in the
adsorption mechanism (dyedye interactions). Despite the number of papers published, there is as yet
little literature containing a full study comparing
various models and this topic clearly needs further
detailed research. Some conclusions can be found in
Refs. [93,97,98].
The second point is related to the mathematical
model used. There is no doubt that mathematical
modeling is an invaluable tool for the analysis and
design of adsorption systems and also for the
theoretical evaluation and interpretation of thermodynamic parameters. However, an isotherm may
t experimental data accurately under one set of
conditions but fail entirely under another. In
addition, no single model has been found to be
generally applicable. This is readily understandable
in the light of the assumptions associated with their
respective deviations. In the single-component isotherm studies, the optimization procedure requires
an error function to be dened in order to
quantitatively compare the applicability of different
models in tting data. To determine isotherm
constants for two-parameter isotherms such as the
Langmuir and the Freundlich models, two methods
are available: tting the isotherm equation to
the data in its non-linear form or converting the
equation into a linear form by transforming the
isotherm variables. In the literature, linear regression is the most commonly used method to estimate
adsorption, and linear coefcients of determination
are preferred. However, the use of this method is
limited to solving linear forms of equation which
measure the difference between experimental data
and theoretical data in linear plots only, but not the

435

errors in isotherm curves. Much work is also


required to this area. Recently, several studies have
shown that the linearization of a non-linear
isotherm expression produces different outcomes
[73,97]. Crini et al. [73] reported that linear
regression and the non-linear Chi-square analysis
gave different models as the best-tting isotherm for
the given data set, thus indicating a signicant
difference between the analytical methods. They
showed that the non-linear Chi-square test provided
a better determination for the experimental data.
Wong et al. [97] also reported that the values of the
individual isotherm constants changed with the
error methodology selected. They obtained contradicting results from linearization using different
error functions.
6.2. Kinetic modeling
An ideal adsorbent for wastewater pollution
control must not only have a large adsorbate
capacity but also a fast adsorption rate. Therefore,
the adsorption rate is another important factor for
the selection of the material and adsorption kinetics
must be taken into account since they explain how
fast the chemical reaction occurs and also provides
information on the factors affecting the reaction
rate. The kinetics of adsorption is also another area
of debate, and once again, differences in chitosan
type, preparation, dyes and methodology examined
makes any comparison of results difcult.
Three kinetic models (Table 10) have been widely
used in the literature for adsorption processes: (i)
pseudo-rst-order kinetic model (Lagergren model)
[127]; (ii) pseudo-second-order kinetic model (Ho
and McKay model) [128]; (iii) and intraparticle
diffusion model (Webber and Morris model) [129].
These kinetic models are used to examine the
controlling mechanism of adsorption process such

Table 10
The three most popular kinetic model and their linear forms
Model
Lagergren

Ho and McKay

Webber and Morris

Equation


qe
k1
t

log
qe  qt
2:303

Linear form

1
1
k2 t
qe  qt qe

t
1
1

t
qt k2 q2e qe

logqe  qt log qe 

qt ki t1=2 C

k1
t
2:303

ARTICLE IN PRESS
436

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

as adsorption surface, chemical reaction and/or


diffusion mechanisms (Table 8). The parameters of
the kinetic models can be obtained by suitable
linearization procedures followed by both linear
and/or non-linear regression analysis. Recent work
by Crini et al. [73] has shown that non-linear
regression gives a more accurate determination of
parameters than linear methods.
When adsorption is preceded by diffusion
through a boundary, the kinetics in most cases
follows the pseudo-rst-order rate equation of
Lagergren. Dutta et al. [75] reported that the
adsorption of reactive and direct dyes on chitosan
followed rst-order kinetics and the Lagergren plots
were linear for a wide range of concentrations and
contact periods. Chang and Juang [86] and Wong
et al. [97] also indicated that the pseudo-rst-order
equation could well describe the adsorption processes. However, the Lagergren model was not
proved to be effective in representing the experimental kinetic data for the entire adsorption period.
In some cases though the Lagergren model provided
an excellent t with the experimental kinetic data, it
failed to predict the amount of dye adsorbed
theoretically thereby deviating from theory. So,
the pseudo-second order was preferred.
The second order (and also the rst order) is
based on the adsorption capacity: it only predicts
the behavior over the whole range of studies
supporting the validity, and is in agreement with
chemisorption being the rate-limiting step. The
kinetics of adsorption of many dye species onto
various chitosan materials was also found to be of
second-order in the literature: adsorption of AO 7
[67], DR 81 [67], RR 222 [69] and RR 189 [68,94] on
crosslinked chitosan, AR 87 [83], RR 2 [67] and RB
222 [90] on raw chitosan (Table 8). The applicability
of the pseudo-second-order model suggested that
chemisorption might be the rate-limiting step that
controls these adsorption processes. In general, this
model is interesting and useful since the Ho and
McKay [128] equation was found to explain the
kinetics of most adsorption systems very well for the
entire range of adsorption periods using different
concentrations and chitosan dosages. In addition, it
has the following advantage: the adsorption capacity, the pseudo-second-order rate constant, and the
initial adsorption rate can be determined from the
equation without knowing any parameters beforehand [133]. Crini et al. [73] used different kinetic
models for the characterization of the adsorption of
BB 9 and BB 3 by grafted chitosan. The kinetic

measurements and their modeling showed that both


processes were rapid because of rapid surface
physical adsorption and the Ho and McKay
equation was more accurate at tting the experimental data. They reported that, at all initial dye
concentrations, the adsorption data were well
represented by the Lagergren model for only the
rst 60 min and thereafter they deviated from
theory. The adsorption data were well represented
only in the region where rapid adsorption took
place. This conrmed that it was not appropriate to
use the Lagergren kinetic model to predict the
adsorption kinetics of BB 3 onto chitosan for the
entire adsorption period. The adsorption system
obeys the pseudo-second-order kinetic model for
the entire adsorption period and thus supports the
assumption behind the model that the adsorption is
due to chemisorption. They also showed that the
kinetic parameters decreased markedly with increasing initial dye adsorption. The adsorption of dye
probably takes place via surface exchange reactions
until the surface functional sites are fully occupied;
thereafter dye molecules diffuse into the polymer
network for further interactions and/or reactions.
Both the Lagergren, and Ho and McKay models
basically include all steps of adsorption (i.e. external
lm diffusion, adsorption and intraparticle diffusion), they are thus pseudo-models [86]. However,
using the so-called pseudo-rst and pseudosecond-order equations for data interpretation is
questionable since the equations have no physical
signicance. It is more reasonable to interpret the
kinetic data in terms of mass transfer [130132].
During the past several decades, a large number of
studies of batch adsorption have been reported in
the literature and a summary of these studies can be
found in the excellent compilation reported by Tien
[122]. Because the above two-lumped kinetic pseudo-models cannot identify adsorption mechanisms,
several investigators proposed to use the diffusion
mechanisms such as intraparticle diffusion using the
Weber and Morris equation, the Avrami model and
the Elovich equation [86]. The former model
originates from Ficks second law. The validity of
the Elovich equation suggests that the chemisorption (chemical reaction) mechanism is probably rate
controlling in the adsorption mechanism. Chang
and Juang [86] indicated that although the pseudorst-order equation could describe the adsorption
processes well, from a lumped point of view, the
better-t of kinetic data by the Elovich equation
instead of by intraparticle diffusion suggested the

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

signicance of chemisorption mechanism during the


processes. They supposed that the coordination and
reaction between the dyes and the amino and
hydroxy groups on chitosan chains would be
signicant and chemisorption controlled the process. Cestari et al. [62,63] indicated that the pseudomodels did not take into account the inuence
of parameters such as temperature and kind of
dye. So, they suggested using the Avrami model,
which is the best kinetic model to evaluate multistep
adsorption phenomena at the solid/solution interface. However, this model cannot give interaction
mechanisms.
Mass transfer involves several steps including (i)
bulk diffusion, (ii) lm diffusion, (iii) intraparticle
diffusion and (iv) (physical and/or chemical) adsorption reactions. Numerous authors consider that
bulk and lm diffusion can be ignored if a sufcient
stirring speed is used. This is correct for bulk
diffusion but is more controversial regarding lm
diffusion. Moreover, it is usually accepted that, in
the case of physical adsorption, the adsorption itself
can be considered as an instantaneous processes,
and the adsorption kinetics are controlled either by
external or intraparticle diffusion or by both
diffusion mechanisms at the same time [122]. In
the case of chemical reactions, their own kinetic
rates may interfere in the control of the adsorption
rate. For complete modeling of adadsorption
kinetics it would be necessary to take into account
not only the diffusion equations but also boundary
conditions including the adsorption isotherm equation [18,122]. This means that the system of
equations is very complex but, generally, it is
possible to simplify the system by separating
diffusion steps or taking into account only diffusion
steps in the control of kinetic rates. In different
adsorption studies, the diffusion mechanisms were
considered independently in accordance with the
assumptions that the kinetics was controlled by
external diffusion at the beginning of the experiment
and then controlled by intraparticle diffusion.
McKay [47] observed that diffusion within the
particle is much slower than the movement of the
dye from solution to the external solid surface
because of (i) the greater mechanical obstruction to
movement presented by the surface molecules or
surface layers and (ii) the restraining chemical
attractions between dye and adsorbent. During
adsorption of the dye from a batch system, dye
molecules arrive at the adsorbent surface more
rapidly than they can diffuse away into the solid.

437

The dye accumulates at the surface and a (pseudo)equilibrium is established, and further adsorption of
dye can take place only at the same rate as the
surface concentration is depleted by inward adsorption. The dye uptake can be correlated to the square
of time over a large adsorption zone to get
diffusivity of the dye in adsorbent particles. In
diffusion studies, it is possible to dene a rate
parameter by plotting the adsorption capacity as a
function of the square root of time [47,93]. The root
time dependence may be expressed by the equation
proposed by Weber and Morris [129], assuming that
the mathematical dependence is obtained if the
process is considered to be inuenced by simple
diffusion in the particles and convective diffusion in
the solution. If intraparticle diffusion is involved in
the adsorption process, then the plot of the square
root of time versus the uptake would result in a
linear relationship, and intraparticle diffusion
would be the rate-limiting step if this line passed
through the origin. When the plots do not pass
through the origin, this is indicative of some degree
of boundary layer control and further shows that
the intraparticle diffusion is not the only ratecontrolling step, but that other processes may
control the rate of adsorption. The WebberMorris
plot is also an empirically relationship but widely
used in the literature.
Several different steps in the process have been
characterized by this simple mathematical model
and different linear sections have been identied in
the adsorption of dyes on chitosan as reported by
numerous authors [59,66,68,73,87,89,91]. All the
studies showed that the kinetics results can be used
to determine if particle diffusion is the rate-limiting
step for dye adsorption onto a material. In general,
the WebberMorris plots present a multilinearity,
which indicates that two or more steps occur in the
process. In the plots, there are three different
portions, representing the different stages in adsorption: an initial curved portion followed by
linear portion and then a plateau. The initial curve
portion is due to surface adsorption and rapid
external diffusion (boundary layer diffusion). The
second linear portion is the gradual adsorption
stage where the intraparticle diffusion is ratecontrolled. The plateau (third portion) is the nal
equilibrium stage, where the intraparticle diffusion
starts to slow down due to the low solute
concentration in solution. Juang and co-workers
[90,91,93] reported that adsorption kinetics were
controlled by different mechanisms, the most

ARTICLE IN PRESS
438

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

limiting of which were the diffusion mechanisms


including the external and the intraparticle mass
transfer resistances and the reaction rate. The
WebberMorris plots gave three-stage sections,
which mean an instantaneous adsorption stage, a
gradual adsorption stage and nal equilibrium stage
in sequence. As a rst approximation, external
diffusion controls the initial stage of adsorption
process while the second stage of the process is
controlled by the intraparticle diffusion. Similar
conclusions have been reported by other authors
[72,73,81,83]. Crini et al. [72,73] reported that the
multilinearity obtained using the WebberMorris
model showed a contribution of lm diffusion on
the control of adsorption kinetics and the intraparticle diffusion played an important role but was not
the rate-determining step. Two diffusion mechanisms are involved in the adsorption rate: pore
diffusion (diffusion within the pore volume) and
surface diffusion (diffusion along the surface of the
pores). Pore diffusion and surface diffusion occur in
parallel within the adsorbent particle. They concluded that the mechanism was complex, involving
adsorption on the external surface, diffusion into
the bulk, chemisorption and other interactions
(mainly hydrophilic and hydrophobic interactions).
Gibbs et al. [81] also observed that the adsorption of
AG 25 on chitosan appeared to occur not only at
the surface of the material but in its intraparticle
network with chemisorption the rate-limiting step.
They concluded that the resistance to intraparticle
diffusion also plays an important role in the control
of mass transfer.
Several factors can affect the reaction kinetics of
dye adsorption onto chitosan. These factors include
the chemical structure of the target dye, the
characteristics of the adsorbent (in particular, its
particle size) and/or the experimental solution
conditions. Guibal and co-workers [8082] reported
that adsorption kinetics were strongly inuenced
not only by intraparticle diffusion resistance but
also by the afnity of the dye for the material. The
afnity of the dye molecule for the adsorbent
changed the relative importance of the intraparticle
diffusion on the control of the overall kinetics. The
concentration of the dye could also strongly affect
the kinetics [80]. The strong effect of particle size
also conrmed that the contribution of intraparticle
diffusion resistance to the control of kinetics cannot
be neglected [82]. The greater the particle size, the
greater the contribution of intraparticle diffusion
resistance to the control of the adsorption kinetics

for only slightly porous materials. They indicated


that the size of adsorbent particles inuenced both
the adsorption kinetics and equilibrium for AG 25
[81] because of resistance to intraparticle diffusion,
but the porosity of the sorbent and its surface area
did not control the adsorption kinetics [82] for
numerous anionic dyes. In the case of RB 5 on
chitosan, they observed that the kinetic parameters
varied little and the most signicant effect observed
was the decrease in intraparticle diffusivity [80].
Juang et al. [93] also reported a greater effect of
particle size on reactive dye adsorption kinetics by
chitosan. These authors indicated that the greater
the amount and the smaller the size of the chitosan
particles used, the faster the process. Wu et al. [91]
found that the adsorption was faster using beadtype chitosan than the ake type.
6.3. Thermochemistry of biosorption
6.3.1. Effect of temperature
Generally speaking, the adsorption of pollutants
increases with temperature because high temperatures provide a faster rate of diffusion of adsorbate
molecules from the solution to the adsorbent [134].
However, it well known that temperature plays an
important role in adsorption in activated carbon,
generally having a negative inuence on the amount
adsorbed. The adsorption of organic compounds
(including dyes) is an exothermic process and the
physical bonding between the organic compounds
and the active sites of the carbon will weaken with
increasing temperature. Also with the increase of
temperature, the solubility of the dye also increases,
the interaction forces between the solute and
the solvent become stronger than those between
solute and adsorbent, consequently the solute is
more difcult to adsorb. Both of these features are
consistent with the order of Langmuir adsorption
capacity. The adsorption of dyes by chitosan is also
usually exothermic: an increase in the temperature
leads to an increase in the dye adsorption rate, but
diminishes total adsorption capacity [21,135]. However, these effects are small and normal wastewater
temperature variations do not signicantly affect
the overall decolorization performance [21]. In
addition, the adsorption process is not usually
operated at high temperature because this would
increase operation costs.
The increase in temperature affects not only the
solubility of the dye molecule (its solubility increases) but also the chemical potential of the

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

material (its potential increases), the potential being


a controlling factor in adsorption. Both effects work
in the same direction causing an increase in the
batch system. In general, this could be conrmed by
the thermodynamic parameters. An increase in
temperature is also followed by an increase in the
diffusivity of the dye molecule, and consequently by
an increase in the adsorption rate if diffusion is the
rate-limiting step. Temperature could also inuence
the desorption step and consequently the reversibility of the adsorption equilibrium. So, the
temperature (and its variation) is an important
factor affecting chitosan adsorption and investigations of this parameter offer interesting results,
albeit often contradictory.
Annadurai [59], studying BB 9 adsorption on
chitosan, found that adsorption increased with
temperature, peaking at 60 1C. Cestari et al. [61]
indicated that the adsorption behavior of anionic
dyes was directly related to the adsorption temperature. They reported that the dimensions of
the chitosan pores increased with temperature. The
greater the particle pore sizes, the smaller the
contribution of intraparticle diffusion resistance.
So, the increase with the temperature seems to
decrease the impact of the boundary-layer effect.
However, they concluded that dependencies in
relation to both the chemical structure of the dye
molecules and the temperature were not clearly
identied. Dutta et al. [75] studying the adsorption
of reactive and direct dyes on chitosan also observed
that as the temperature of the solution increased so
did the extent of adsorption. Uzun and Guzel [114]
reported that the adsorption of Rb 5 by chitosan
and O II by grafted chitosan must be studied at high
temperatures. They explained their results on the
basis of strong chemical adsorption since the dyes
were more reactive at higher temperatures. In
another recent work [112], they also indicated that
the adsorption of RY 2 must be studied at high
temperatures. These authors concluded that the
adsorption capacity of chitosan strongly increased
with increasing temperature. The observed increase
in adsorption may be attributed to the fact that on
increasing temperature, a greater number of active
sites is generated on the polymer beads because of
an enhanced rate of protonation/deprotonation of
the functional groups on the beads. The fact that
adsorption of dyes on chitosan increases with higher
temperature can be surprising. Temperature is well
known to play an important role on adsorption in
activated carbon, generally having a negative

439

inuence on the amounts adsorbed. The adsorption


of organic compounds (including dyes) is an
exothermic process (negative value of enthalpy
change) which is responsible for reduction in
adsorption as the temperature is increased. As
mentioned above, the physical bonding between
the organic compounds and the active sites
of the carbon will weaken with increasing temperature. The fact that an increase in temperature is
followed by a decrease in adsorption capacity
suggests that adsorption is governed only by
physical phenomena. Also with the increase of
temperature, the solubility of the dyes also increases, the interaction forces between the solute
and the solvent become stronger than solute and
adsorbent, consequently the solutes are more
difcult to adsorb. Both of these features are
consistent with the order of Langmuir adsorption
capacity.
Other authors concluded that an increase in
temperature leads to a decrease in the amount of
adsorbed dye at equilibrium since adsorption on
chitosan is exothermic. Saha et al. [102], studying
the adsorption of an azo dye onto chitosan akes
noted that the adsorption capacity was remarkably
reduced with increasing solution temperature. They
concluded that the decrease of the equilibrium
uptake with the increase in temperature means that
the dye biosorption process is exothermic. Li et al.
[94] reported that the adsorption of RR 189 on
crosslinked chitosan was slightly inuenced by
temperature. Thermodynamic parameters such as
the Gibbs free energy change (DG) or enthalpy
change (DH), and/or the apparent activation energy
(Ea) are often used for the characterization of the
temperature effect. For example, more negative
values of DG at higher temperatures imply the
greater driving force of adsorption at high temperatures than at low. The magnitude of activation
energy gives the type of adsorption, which is mainly
physical (physisorption) or chemical (chemisorption). The range of 540 kJ/mol of activation
energies corresponds to a physisorption mechanism
and the range of 40800 kJ/mol suggests a chemisorption mechanism. The values of Ea obtained in
two previous studies [68,94] indicated that the
adsorption on crosslinked chitosan had a low
potential barrier which was assigned to physisorption. These values of heat of adsorption were
estimated from the integrated vant Hoff equation,
which relates the Langmuir equilibrium constant to
the temperature.

ARTICLE IN PRESS
440

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

6.3.2. Thermodynamic parameters


The most important features involved in the
investigation of adsorption phenomenon are
the adsorption isotherm and kinetics, the interface
characteristics, the adsorbateadsorbent interactions, and also the thermochemistry of adsorption.
In particular, the adsorption characteristics of a
material can be expressed in thermodynamic parameters such as DG (Gibbs free energy change),
DH (enthalpy change), and DS (entropy change).
These parameters can be calculated by using the
thermodynamic equilibrium coefcient obtained
at different temperatures and concentrations. The
expressions reported in Table 11 are used. Evaluation of these parameters gives an insight into the
possible mechanisms of adsorption.
The original concepts of thermodynamics assumed that in an isolated system where energy
cannot be gained or lost, the entropy change is the
driving force. In environmental engineering practice, both energy and entropy factors must be
considered in order to determine what processes
will occur spontaneously. Thermodynamic considerations tell us that, at constant temperature and
pressure, the DG value is the fundamental criterion
of spontaneity, and a negative value for DG stands
for the adsorption to take place, indicating the
spontaneity of the reaction. By using the equilibrium constant obtained for each temperature from
the Langmuir model, DG can be calculated according to the Gibbs expression (Table 11). It is
important to note that DG is estimated from the
equilibrium adsorption data under the assumption
that the adsorption of a molecule is reversible and
that an equilibrium condition is established in the
batch system.

The DH and DS changes of an adsorption


reaction can be determined using the vant Hoff
plot (Table 11) and are estimated by determining the
isotherm at different temperatures assuming these
parameters to be independent of temperature. From
a more random stage (in solution) to a more orderly
stage (on the surface of the adsorbent) for dye
molecules, the entropy change of adsorption (DS)
also has a negative sign. The sign of DS would
indicate the direction, for adsorption (+DS), and
for desorption (DS). As known from thermodynamics, the negative values of DG and DS require a
negative adsorption enthalpy (DH), which in turn
implies that the adsorption phenomenon is exothermic. The DH value (experimentally measured) can
also be used as a measure of the interaction force
between adsorbate and adsorbent, giving an indication of the bonding strength.
Adsorption on solids is classied into physical
adsorption and chemical adsorption, but the dividing line between the two is not sharp. However,
physical adsorption is non-specic, and the variation of energy for physical adsorption is usually
substantially smaller than that of chemical adsorption. Chemical adsorption is similar to ordinary
chemical reactions in that it is highly specic.
Typically, DH for physical adsorption ranges from
4 to 40 kJ/mol, compared to that of chemical
adsorption ranging from 40 to 800 kJ/mol. As
shown in Table 12, the DH values suggest that the
adsorption process might be considered as physical
adsorption in nature. Table 12 also shows that, for
dye adsorption on chitosan derivatives, negative DG
values reveal the spontaneity of the process. The
adsorption process is spontaneous in nature and
more favorable at lower concentrations of dye

Table 11
Thermodynamic equations and their parameters
Expression

Linear equation form

Arrhenius

ln kads 

Gibbs

DG RT ln K L

vant Hoff

ln K L 

Clausius-Clapeyron

RT 1 T 2
ln C 2  ln C 1
DH
T2  T1

Parameters

Ea
ln ko
RT

DH DS

RT
R

Apparent activation energy

Free energy change


with K L qe =C e

Enthalpy change

Entropy change
Enthalpy change

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

441

Table 12
Thermodynamics and rate parameters for various dyes using chitosan
Dye

Material

T (1C)

pH

AB
AR 87
Azo
BB
BB 1
BB 3
BB9
BV 3
CV
CV
CV
IC
IC
O II
RR 141
RR 189
RR 189

Chitosan

20
30
30
20
30
25
25
30
20
20
60
35
25
20
20
30
30

3.6

Chitosan
Grafted chitosan
Grafted chitosan
Grafted chitosan
Grafted chitosan
Chitosan
Modied chitosan
Modied chitosan

Chitosan
Crosslinked bead
Crosslinked bead

DG (kJ/mol)

5.46
4.97
6.9
3

11
3
3

11.89
6.4
22.1
11.67
10.51
11.72
10.17
9.1
2.55
3.35
7.10
6.6

DH (kJ/mol)
72.87
17.10
2.17
47.46

DS (J/molK)

k0 (kg/g min)

Ea (kJ/mol)

Reference

75.7
43

[65]
[83]
[102]
[65]
[106]
[73]
[58]
[106]
[114]
[114]
[114]
[63]
[101]
[114]
[109]
[68]
[94]

51.8
9.25

2.47
1.09
23.05
23.05
23.2
29.25
5.68
18.20
52.9

compared with higher concentrations, as reported


by Chatterjee et al. [83]. The appreciably low free
energy values indicated saturation of the process
and the enthalpy values suggested that the reaction
was exothermic, and especially favorable at low
temperature. DG was more negative with decreasing
temperature, which suggested that lower temperature makes the adsorption easier, as observed by
Uzun and Guzel [114]. However, the authors [114]
concluded that the dye adsorption by chitosan must
be studied at high temperatures. Saha et al. [102],
studying the adsorption of an azo dye onto chitosan
akes, also reported that the Gibbs free energy
demonstrated that the adsorption was favorable
and the pronounced chitosandye interaction was
reected in the values of enthalpy. Prado et al. [100]
concluded similar observations. Indigo carmine/
chitosan interaction showed favorable enthalpic
and entropic processes, reecting thermodynamic
stability of the complex formed (Scheme 1), while
the dye/chitin interaction showed an exothermic
enthalpy and a highly unfavorable entropic effect,
resulting in a non-spontaneous thermodynamic
system. Other observations were also found: the
positive values of DH [58] and DS [102,110,114]
suggested the endothermic nature of adsorption and
increased randomness at the solid/solution interface
during the adsorption of dye on chitosan derivatives. A low value of DS indicated that no
remarkable changes in entropy occur [114]. Chen
et al. [65], studying the adsorption of acid (AB) and
basic (BB) dyes on chitosan reported that the

32.14
38.66
38.65
45.8
90
7.94
37.88
153.1

8.11  107

negative values of DH for AB dye indicated that


heat was released during the adsorption process and
the positive value for BB indicated that heat was
abstracted from the surroundings They concluded
that the effect of temperature on the adsorption of
cationic dye was peculiar. There is an sharp increase
in equilibrium adsorption with increased temperature, which was thought to be due to enhanced dye
mobility and a temperature-induced swelling effect
within the internal structure of the chitosan,
allowing the large dye ions to penetrate into the
particles. In addition, for AB, the adsorption rate
and intraparticle diffusion coefcient were much
larger than for BB. In other words, the intraparticle
diffusion of AB was more rapid. The diffusion
coefcient decreased with increasing temperature.
This could be explained by the fact that with
increasing temperature, the amount of dye adsorbed
on the active sites increased in the early stages,
leading to a decrease in the mobility of the diffusion
molecules to pass through for adsorption in the
long-term stage. Thus, the diffusion rate of dyes in
the intraparticle diffusion process decreased with
increasing temperature.
In general, in the external mass transport process,
the values of the diffusion coefcient increase as the
temperature of adsorption increases. When the
temperature increases, the thickness of the boundary layer surrounding the adsorbent and the mass
transport resistance of the adsorbate in the boundary layer decreases. Thus, the diffusion rate of dyes
in the external mass transport process increases with

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

442

temperature, while in intraparticle diffusion, the


coefcient of diffusion values decrease with increasing temperature. At low temperature, the diffusion
coefcient of external mass transport is slightly
lower than the diffusion coefcient of intraparticle
diffusion. So, at low temperature, dye adsorption is
limited by the external mass transport. With the
increase of temperature external mass transport
begins to play a major role in dye adsorption by
chitosan. One of the reasons for the positive changes
of the enthalpy and entropy could be the release of
numerous water molecules. The adsorption of the
hydrated (poly)anions onto a hydrophilic polymer
network inevitably disturbs the order of water
molecules in the nearest environment and releases
them to the external liquid. In other words,
adsorbed molecules are attracted probably due to
long-distance electrostatic interactions between oppositely charge groups. During the formation of the
ionic bonds between the dye and the polymer, the
counterions should gain a higher degree of freedom
and increase the entropy.
7. Economic aspects
Research is mainly focused on the technical
performances of chitosan derivatives, while their
economic aspect is usually neglected. Cost is
actually an important parameter for comparing
adsorbent materials. According to Bailey et al.
[136], a sorbent can be considered low cost if it
requires little processing, is abundant in nature, or is
a by-product or waste material from another
industry. Chitosan-based materials display economic advantages:

Chitin is a material obtained from natural raw


resources. It is only commercially extracted from
crustaceans which are conveniently available as
waste from processing shellsh. The wastes
consists of chitin (2030%), proteins (2040%),
salts (mainly carbonate and phosphate, 3060%)
and lipids (014%) [35]. These proportions vary
with species and season. Several countries possess
large unexploited crustacean resources, especially
in Asia.
Chitin and chitosan are now produced commercially at low cost and their production is also
economically interesting, especially if it includes
the recovery of carotenoids. A prerequisite for
the greater use of chitin in industry is cheap
manufacturing processes and/or the development

of protable processes to recover chitin and byproducts such as proteins and pigments. It is
important to note that the recovery of these
products from waste is an additional source of
revenue Crustacean shells contains considerable
quantities of carotenoids which so far have not
been synthesized, and which are marketed as a
sh food additive in aquaculture, mainly for
salmon [21]. In addition, calcium carbonate
which is another major component of crab shells,
is converted to calcium oxide and sodium
carbonate [12]. Pigments may be also recovered
as high value side products.
The production of the chitosan-based materials is
economically feasible because they are easy to
prepare with relatively inexpensive chemical
reagents under mild conditions. The procedures
also require relatively harmless chemicals.

However, the industrial isolation of the polymers


is restricted due to problems of environmental
pollution. The traditional method of extraction
creates its own environmental problems as it
generates large quantities of concentrated efuent
containing polluting bases and degradation products and presenting inconsistent physicochemical
properties. At the same time, the conversion to
chitosan at high temperature with strong alkali can
cause variability of product properties and chitosan
quality, and can also increase the processing costs.
This also appears to have limited potential for
industrial acceptance. Recently, some other sources
such as yeast and fungi (zygomycetes) have begun to
be employed to obtain chitosan. They can be readily
cultured in simple nutrients and used as an
alternative source of chitosan. With advances in
fermentation technology chitosan preparation from
fungal cell walls will become an alternative route for
the production of this polymer via an eco-friendly
pathway.
8. Concluding remarks
The state-of-the-art in the eld of biosorption of
dyes by chitosan using batch systems is reviewed in
this paper, based on a substantial number of
relevant references published recently. Of course,
this is an ambitious project since a direct comparison of data obtained using different materials is
difcult to make. The experimental conditions used
in the batch system are also not systematically the

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

same. Nevertheless, the following conclusions may


be reached:

The works reviewed above indicate that bioadsorption onto chitosan is becoming a promising
alternative to replace conventional adsorbents
used for decolorization purposes. Outstanding
progress has been made, demonstrating the
application of chitosan and crosslinked chitosan
in dye bioadsorption. These materials are
efcient in dye removal with the additional
advantage of being cheap, non-toxic and biocompatible.
There is abundant literature concerning the
evaluation of adsorption performances of chitosan, especially in terms of adsorption capacity
(amount of dye adsorbed). At least 100 dyes
have been studied so far. All the studies showed
that chitosan had an extremely high afnity for
many classes of dyes. In particular, it has
demonstrated outstanding removal capacities
for acid, reactive and direct dyes. However,
dependencies in relation to the chemical structure
of the dyes were not clearly identied and there
is, as yet, little information in the literature on
this topic.
Chitosan is characterized by its easy dissolution
in many dilute mineral acids, with the remarkable
exception of sulfuric acid. It is thus necessary to
stabilize it chemically for the recovery of dyes in
acidic solutions. Several methods have been
developed to reinforce chitosan stability. The
advantage of chitosan over other polysaccharides
is that its polymeric structure allows specic
modications without too many difculties. The
chemical derivatization of the polymer by grafting new functional groups onto the chitosan
backbone may be used to increase the adsorption
efciency, to improve adsorption selectivity, and
also to decrease the sensitivity of adsorption
environmental conditions.
It is interesting to note the relationships between
physicochemical properties and/or sources of
chitosan and the dye-binding properties. Most
of the properties and potential of chitosan as
adsorbent can be related to its cationic nature,
which is unique among abundant polysaccharides and natural polymers, and its high charge
density in solution.
While adsorption on activated carbons is largely
independent of the pH, the adsorption of dyes on
chitosan is controlled by the acidity of the

443

solution in the case of anionic dyes. It is


important to indicate that a source of discrepancies in published studies may be related to
misunderstanding the impact of pH variation
on the adsorption performance.
However, which adsorbent is better: chitosan
(raw material, preconditioned chitosan, grafted
or crosslinked chitosans) or CAC? There is no
direct answer to this question because the best
choice depends on the dye and it is impossible to
determine a correlation between the chemical
structure of the dye and its afnity for either
carbon or chitosan. Each product has advantages
and drawbacks. In addition, comparisons are
difcult because of the scarcity of information
and also inconsistencies in data presentation.

Although extensive work has been done, future


research needs to look into some of the following
aspects:

The biosorbent and the dye structure: It is


necessary to continue to search for and select
the most promising types of chitosan. To date,
there is no systematic and comparative study
taking into account the physicochemical properties of the different kind of dyes. A more detailed
study appears to be necessary to show how the
chemical structure of the dyes affects not only the
adsorption capacities but also the understanding
of adsorption phenomenon involved in the
uptake of a given dye. Recently, some investigators have focused on studying the inuence of the
chemical structure of dyes on adsorption capacity. These studies would help in optimizing the
type and amount of chitosan. The development
of mechanistic and mathematical models in order
to simulate the adsorption process and to
characterize the interaction between the surface
of the chitosan and the adsorbed species are also
important aspects in future biosorption studies,
and should be developed.
Real effluent: The experimental conditions should
be chosen to simulate real wastewater on the
basis of thermodynamics and reaction kinetics
studies;
Large-scale experiment: Biosorption processes
are basically at the stage of laboratory-scale
study in spite of unquestionable progress. Much
work in this area is necessary to demonstrate the
possibilities on an industrial scale.

ARTICLE IN PRESS
444

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447

Acknowledgements
The authors wish to thank Nadia Morin-Crini
and Brigitte Jolibois (LBE, University of FrancheComte) and gratefully acknowledge the nancial
support of the OSEO ANVAR of Franche-Comte.

References
[1] Ali M, Sreekrishnan TR. Aquatic toxicity from pulp and
paper mill efuents: a review. Adv Environ Res 2001;5:
17596.
[2] Pokhrel D, Viraraghavan T. Treatment of pulp and paper
mill wastewatera review. Sci Total Environ 2004;333:
3758.
[3] Thompson G, Swain J, Kay M, Forster CF. The treatment
of pulp and paper mill efuent: a review. Bioresour Technol
2001;77:27586.
[4] Robinson T, McMullan G, Marchant R, Nigam P.
Remediation of dyes in textile efuent: a critical review
on current treatment technologies with a proposed alternative. Bioresour Technol 2001;77:24755.
[5] Banat IM, Nigam P, Singh D, Marchant R. Microbial
decolorization of textile-dye-containing efuents: a review.
Bioresour Technol 1996;58:21727.
[6] Crini G. Non-conventional low-cost adsorbents for dye
removal: a review. Bioresour Technol 2006;60:6775.
[7] Aksu Z. Application of biosorption for the removal of
organic pollutants: a review. Process Biochem 2005;40:
9971026.
[8] Forgacs E, Cserhati T, Oros G. Removal of synthetic dyes
from wastewaters: a review. Environ Int 2004;30:95371.
[9] Synowiecki J, Al-Khateeb NA. Production, properties, and
some new applications of chitin and its derivatives. Crit
Rev Food Sci Nutrition 2003;43:14571.
[10] Struszczyk MH. Chitin and chitosanpart I. Properties
and production. Polimery 2002;47:31625.
[11] Rinaudo M. Chitin and chitosan: properties and applications. Prog Polym Sci 2006;31:60332.
[12] Peter MG. Application and environmental aspects of chitin
and chitosan. J M S Pure Appl Chem 1995;32:62940.
[13] Cheng WP, Chi FH, Yu RF, Lee YC. Using chitosan as a
coagulant in recovery of organic matters from the mash
and lauter wastewater of brewery. J Polym Environ 2005;
13:3838.
[14] Meyssami B, Kasaeian AB. Use of coagulants in treatment
of olive oil wastewater model solutions by induced air
otation. Bioresour Technol 2005;96:3037.
[15] Huang C, Chen S, Pan JR. Optimal condition for
modication of chitosan: a biopolymer for coagulation of
colloidal particles. Water Res 2000;34:105762.
[16] Franks GV. Stimulant sensitive occulation and consolidation for improved solid/liquid separation. J Colloid Int Sci
2005;292:598603.
[17] Chen L, Chen D, Wu C. A new approach for the
occulation mechanism of chitosan. J Polym Environ 2003;
11:8792.
[18] Guibal E. Interactions of metal ions with chitosan-based
adsorbents: a review. Sep Purif Technol 2004;38:4374.

[19] Varma AJ, Deshpande SV, Kennedy JF. Metal complexation by chitosan and its derivatives: a review. Carbohydr
Polym 2004;55:7793.
[20] Struszczyk MH. Chitin and chitosanpart II. Applications
of chitosan. Polimery 2002;47:396403.
[21] Ravi Kumar MNV. A review of chitin and chitosan
applications. React Funct Polym 2000;46:127.
[22] No HK, Meyers SP. Application of chitosan for treatment
of wastewaters. Rev Environ Contam Toxicol 2000;163:
128.
[23] Jayakumar R, Prabaharan M, Reis RL, Mano JF. Graft
copolymerized chitosanpresent and status applications.
Carbohydr Polym 2002;62:14258.
[24] Muzzarelli RAA, Muzzarelli C. Chitosan chemistry:
relevance to the biomedical sciences. Adv Polym Sci 2005;
186:151209.
[25] Sashiwa H, Aiba S. Chemically modied chitin and
chitosan as biomaterials. Prog Polym Sci 2004;29:
887908.
[26] Berger J, Reist M, Mayer JM, Felt O, Peppas NA, Gurny
R. Structure and interactions in covalently and ionically
crosslinked chitosan hydrogels for biomedical applications.
Eur J Pharm Biopharm 2004;57:1934.
[27] Kato Y, Onishi H, Machida Y. N-succinyl-chitosan as a
drug carrier: water-insoluble and water-soluble conjugates.
Biomaterials 2004;25:90715.
[28] Khor E, Lim LY. Implantable applications of chitin and
chitosan. Biomaterials 2003;24:233949.
[29] Senel S, McClure SJ. Potential applications of chitosan in
veterinary medicine. Adv Drug Delivery Rev 2004;56:
146780.
[30] Krajewska B. Applications of chitin- and chitosan-based
materials for enzyme immobilizations. Enzyme Microbial
Technol 2004;35:12639.
[31] Macquarrie DJ, Hardy JJE. Applications of functionalised
chitosane in catalysis. Ind Eng Chem Res 2005;44:
8499520.
[32] Guibal E. Heterogeneous catalysis on chitosan-based
materials: a review. Prog Polym Sci 2005;30:71109.
[33] Krajewska B. Membrane-based processes with use of
chitin/chitosan materials. Sep Purif Technol 2005;41:
30512.
[34] Cagri A, Ustunol Z, Ryser ET. Antimicrobial edible lms
and coatings. J Food Protect 2004;67:83348.
[35] Agullo E, Rodriguez MS, Ramos V, Albertengo L. Present
and future role of chitin and chitosan in food. Macromol
Biosci 2003;3:52130.
[36] Prabaharan M, Mano JF. Chitosan-based particles as
controlled drug delivery systems. Drug Deliv 2005;12:
4157.
[37] Ravi Kumar MNV, Muzzarelli RAA, Muzzarelli C,
Sashiwa H, Domb AJ. Chitosan chemistry and pharmaceutical perspectives. Chem Rev 2004;104:601784.
[38] Giri Dev VR, Neelakandan R, Sudha S, Shamugasundram
OL, Nadaraj RN. Chitosana polymer with wider
applications. Textile Mag 2005;46:836.
[39] Bornet A, Teissedre PL. Applications and interest of chitin,
chitosan and their derivatives in enology. J Int Sci Vigne
et Vin 2005;39:199207.
[40] Mishra G, Tripathy M. A critical review of the treatment
for decolorization of textile efuent. Colourage 1993;40:
358.

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
[41] Prabaharan M, Mano JF. Chitosan derivatives bearing
cyclodextrin cavities as novel adsorbent matrices. Carbohydr Polym 2006;63:15366.
[42] Crini G. Recent developments in polysaccharide-based
materials used as adsorbents in wastewater treatment. Prog
Polym Sci 2005;30:3870.
[43] Giles CH, Hassan ASA, Subramanian RVR. Adsorption of
organic surfaces. IV. Adsorption of sulphonated azo dyes
by chitin from aqueous solutions. J Soc Dyers Colour
1958;74:6828.
[44] McKay G, Blair HS, Gardner JR. Adsorption of dyes on
chitin. I. Equilibrium studies. J Appl Polym Sci 1982;27:
304357.
[45] McKay G, Blair HS, Gardner JR. The adsorption of dyes
on chitin. III. Intraparticle diffusion processes. J Appl
Polym Sci 1983;28:176778.
[46] McKay G, Blair HS, Gardner JR. The adsorption of dyes
onto chitin in xed bed columns and batch adsorbers.
J Appl Polym Sci 1984;29:1499514.
[47] McKay G, Blair HS, Gardner JR, McConvey IF. Tworesistance mass transport model for the adsorption of
various dyestuffs onto chitin. J Appl Polym Sci 1985;30:
432535.
[48] McKay G, Blair HS, Gardner JR. Two-resistance mass
transport model for the adsorption of acid dye onto chitin
in xed beds. J Appl Polym Sci 1987;33:124957.
[49] Yamamoto H. Chiral interaction of chitosan with azo dyes.
Makromol Chem 1984;185:161321.
[50] Maghami GG, Roberts GA. Studies on the interaction of
anionic dyes on chitosan. Makromol Chem 1988;189:
223943.
[51] Seo T, Hagura S, Kanbara T, Iijima T. Interaction of dyes
with chitosan derivatives. J Appl Polym Sci 1989;37:
301127.
[52] Smith B, Koonce T, Hudson S. Decolorizing dye
wastewater using chitosan. Am Dyest Rep 1993;82:
1836.
[53] Stefancich S, Delben F, Muzzarelli RAA. Interactions of
soluble chitosans with dyes in water. I. Optical evidence.
Carbohydr Polym 1994;24:1723.
[54] Delben F, Gabrielli P, Muzzarelli RAA, Stefancich S.
Interactions of soluble chitosans with dyes in water. II.
Thermodynamic data. Carbohydr Polym 1994;24:2530.
[55] Shimizu Y, Kono K, Kim IS, Takagishi T. Effects of added
metal ions on the interaction of chitin and partially
deacetylated chitin with an azo dye carrying hydroxyl
groups. J Appl Polym Sci 1995;55:25561.
[56] Park RD, Cho YY, Kim KY, Bom HS, Oh CS, Lee HC.
Adsorption of toluidine Blue O onto chitosan. Agric Chem
Biotechnol 1995;38:44752.
[57] Park RD, Cho YY, La YG, Kim CS. Application of
chitosan as an adsorbent of dyes in wastewater from
dyeworks. Agric Chem Biotechnol 1995;38:4524.
[58] Lima IS, Ribeiro ES, Airoldi C. The use of chemically
modied chitosan with succinic anhydride in the methylene
blue adsorption. Quim Nova 2006;29:5016.
[59] Annadurai G. Adsorption of basic dye on strongly
chelating polymer: batch kinetics studies. Iran Polym J
2002;11:23744.
[60] Annadurai G. Design of optimum response surface
experiments for adsorption of direct dye on chitosan.
Bioproc Eng 2000;23:4515.

445

[61] Cestari AR, Vieira EFS, Pinto AA, Lopes ECN. Multiple
adsorption of anionic dyes on silica/chitosan hybrid 1.
Comparative kinetic data from liquid- and solid-phase
models. J Colloid Int Sci 2005;292:36372.
[62] Cestari AR, Vieira EFS, dos Santos AGP, Mota JA, de
Almeida VP. Adsorption on of anionic dyes on chitosan
beads. 1. The inuence of the chemical structures of dyes
and temperature on the adsorption kinetics. J Colloid Int
Sci 2004;280:3806.
[63] Dos Anjos FSC, Vieira EFS, Cestari AR. Interaction of
indigo carmine dye with chitosan evaluated by adsorption
and thermochemical data. J Colloid Int Sci 2002;253:
2436.
[64] Chang YC, Chen DH. Adsorption kinetics and thermodynamics of acid dyes on a carboxymethylated chitosanconjugated magnetic nano-adsorbent. Macromol Biosci
2005;5:25461.
[65] Chen L, Chen DH, Gao L. Kinetics of dyes adsorbed by
chitosan. J Dong Hua Univ (Eng Ed) 2002;19:803.
[66] Chiou MS, Chuang GS. Competitive adsorption of dye
metanil yellow and RB15 in acid solutions on chemically
cross-linked chitosan beads. Chemosphere 2006;62:73140.
[67] Chiou MS, Ho PY, Li HY. Adsorption of anionic dye in
acid solutions using chemically cross-linked chitosan beads.
Dyes Pigm 2004;60:6984.
[68] Chiou MS, Li HY. Adsorption behavior of reactive dye in
aqueous solution on chemical cross-linked chitosan beads.
Chemosphere 2003;50:1095105.
[69] Chiou MS, Kuo WS, Li HY. Removal of reactive dye from
wastewater by adsorption using ECH cross-linked chitosan
beads as medium. J Environ Sci Health A Toxic/Hazard
Substances Environ Eng 2003;38:262131.
[70] Chiou MS, Ho PY, Li HY. Adsorption behavior of dye
AAVN and RB4 in acid solutions on chemically crosslinked chitosan beads. J Chin Inst Chem Eng 2003;34:
62534.
[71] Kim TY, Cho SY. Adsorption equilibria of reactive dye
onto highly polyaminated porous chitosan beads. Korean J
Chem Eng 2005;22:6916.
[72] Crini G, Martel B, Torri G. Adsorption of C.I. Basic Blue 9
on chitosan-based materials. Int J Environ Pollut 2008;
33(14) in press.
[73] Crini G, Robert C, Gimbert F, Martel B, Adam O, De
Giorgi F, et al. The removal of Basic Blue 3 from aqueous
solutions by chitosan-based adsorbent: batch studies.
J Hazard Mater doi:10.1016/j.jhazmat.2007.08.025.
[74] Kimura IY, Laranjeira MCM, de Favere VT, Furlan L.
The interaction between reactive dye containing vinylsulfone group and chitosan microspheres. Int J Polym Mater
2002;51:75968.
[75] Dutta PK, Durga Bhavani K, Sharma N. Adsorption for
dyehouse efuent by low cost adsorbent (chitosan). Asian
Textile J 2001;10:5763.
[76] El-Tahlawy KF, Gaffar MA, El-Rae S. Novel method for
preparation of b-cyclodextrin/grafted chitosan and its
application. Carbohydr Polym 2006;63:38592.
[77] Gaffar MA, El-Rae SM, El-Tahlawy KF. Preparation
and utilization of ionic exchange resin via graft copolymerization of b-CD itaconate with chitosan. Carbohydr Polym
2005;56:38796.
[78] Fahmy HM, Mohamed ZE, Abo-Shosha MH, Ibrahim
NA. Thermosole cross-linking of chitosan and utilization in

ARTICLE IN PRESS
446

[79]

[80]

[81]

[82]

[83]

[84]

[85]

[86]

[87]

[88]

[89]

[90]

[91]

[92]

[93]

[94]

[95]

[96]

G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447


the removal of some dyes from aqueous solutions. Polym
Plastics Technol Eng 2004;43:44562.
Guibal E, Touraud E, Roussy J. Chitosan interactions with
metal ions and dyes: dissolved-state vs. solid-state application. World J Microbiol Biotechnol 2005;21:91320.
Gibbs G, Tobin JM, Guibal E. Inuence of chitosan
preprotonation on reactive black 5 adsorption isotherms
and kinetics. Ind Eng Chem Res 2004;43:111.
Gibbs G, Tobin JM, Guibal E. Adsorption of acid green 25
on chitosan: inuence of experimental parameters on
uptake kinetics and adsorption isotherms. J Appl Polym
Sci 2003;90:107380.
Guibal E, McCarrick P, Tobin JM. Comparison of the
adsorption of anionic dyes on activated carbon and
chitosan derivatives from dilute solutions. Sep Sci Technol
2003;38:304973.
Chatterjee S, Chatterjee S, Chatterjee BP, Das AR, Guha
AK. Adsorption of a model anionic dye, eosin Y, from
aqueous solution by chitosan hydrobeads. J Colloid Int Sci
2005;288:305.
Hebeish A, Rafei R, El-Shafei A. The crosslinking of
chitosan with glutaraldehyde for the removal of dyes and
heavy metals ions from aqueous solutions. Tinctoria
2004;101:2834.
Hebeish A, Rafei R, El-Shafei A. Crosslinking of chitosan
with glutaraldehyde for removal of dyes and heavy metals
ions from aqueous solutions. Egypt J Chem 2004;47:6579.
Chang MY, Juang RS. Equilibrium and kinetic studies on
the adsorption of surfactant, organic acids and dyes from
water onto natural biopolymers. Colloid Surf A: Physicochem Eng Aspects 2005;269:3546.
Chang MY, Juang RS. Adsorption of tannic acid, humic
acid, and dyes from water using the composite of chitosan
and activated clay. J Colloids Int Sci 2004;278:1825.
Juang RS, Wu FC, Tseng RL. Use of chemically modied
chitosan beads for adsorption and enzyme immobilization.
Adv Environ Res 2002;6:1717.
Wu FC, Tseng RL, Juang RS. Kinetic modeling of liquidphase adsorption of reactive dyes and metal ions on
chitosan. Water Res 2001;35:6138.
Wu FC, Tseng RL, Juang RS. Enhanced abilities of highly
swollen chitosan beads for color removal and tyrosinase
immobilization. J Hazard Mater 2001;B81:16777.
Wu FC, Tseng RL, Juang RS. Comparative adsorption of
metal and dye on ake- and bead-types of chitosans
prepared from shery wastes. J Hazard Mater 2000;B73:
6375.
Annadurai G, Lee DJ, Juang RS. BoxBehnken studies on
dye removal from water using chitosan and activated
carbon adsorbents. J Chin Inst Chem Eng 2000;31:60915.
Juang RS, Tseng RL, Wu FC, Lee SH. Adsorption
behavior of reactive dyes from aqueous solutions on
chitosan. J Chem Tech Biotechnol 1997;70:3919.
Chiou MS, Li HY. Equilibrium and kinetic modeling of
adsorption of reactive dye on cross-linked chitosan beads.
J Hazard Mater 2002;B93:23348.
Martel B, Devassine M, Crini G, Weltrowski M, Bourdonneau M, Morcellet M. Preparation and adsorption
properties of a beta-cyclodextrin-linked chitosan derivative.
J Polym Sci A: Polym Chem 2001;39:16976.
Paneva D, Stoilova O, Manolova N. Magnetic hydrogel
beads based on chitosan. e-Polymers 2004;60:111.

[97] Wong YC, Szeto YS, Cheung WH, McKay G. Adsorption


of acid dye on chitosanequilibrium isotherm analyses.
Proc Biochem 2004;39:693702.
[98] Wong YC, Szeto YS, Cheung WH, McKay G. Pseudo-rstorder kinetic studies of the adsorption of acid dyes onto
chitosan. J Appl Polym Sci 2004;92:163345.
[99] Wong YC, Szeto YS, Cheung WH, McKay G. Equilibrium
studies for acid dye adsorption onto chitosan. Langmuir
2003;19:788894.
[100] Miyata K, Jin C, Maekawa M. Adsorption removal of
anionic dyes in waste water from dyeing by chitosan. SenI
Gakkaishi 2002;58:4625.
[101] Prado AGS, Torres JD, Faria EA, Dias SCL. Comparative
adsorption studies of indigo carmine dye on chitin and
chitosan. J Colloid Int Sci 2004;277:437.
[102] Saha TK, Karmaker S, Ichikawa H, Fukumori Y.
Mechanisms and kinetics of trisodium 2-hydroxy-1,
1-azonaphthalene-3,4,6-trisulfonate adsorption onto chitosan. J Colloid Int Sci 2005;286:4339.
[103] Shimizu Y, Tanigawa S, Saito Y, Nakamura T. Synthesis
of chemically modied chitosans with a higher fatty acid
glycidyl and their adsorption abilities for anionic and
cationic dyes. J Appl Polym Sci 2005;96:24238.
[104] Shimizu Y, Tominaga T, Saito Y. Diethylaminoethylation
of chitin and the adsorption of acid dyes onto the resulting
polymer. Adsorpt Sci Technol 2004;22:42737.
[105] Shimizu Y, Taga A, Yamaoka H. Synthesis of novel
crosslinked chitosans with a higher fatty diacid diglycidyl
and their adsorption abilities toward acid dyes. Adsorpt Sci
Technol 2003;21:43949.
[106] Chao AC, Shyu SS, Lin YC, Mi FL. Enzymatic grafting of
carboxyl groups on to chitosanto confer on chitosan the
property of a cationic dye adsorbent. Bioresour Technol
2004;91:15762.
[107] Trung TS, Thein-Han WW, Qui NT, Ng CH, Stevens WF.
Functional characteristics of shrimp chitosan and its
membranes as affected by the degree of deacetylation.
Bioresour Technol 2006;97:65963.
[108] Trung TS, Ng CH, Stevens WF. Characterization of
decrystallized chitosan and its application in biosorption
of textile dyes. Biotechnol Lett 2003;25:118590.
[109] Sakkayawong N, Thiravetyan P, Nakbanpote W. Adsorption mechanism of synthetic reactive dye wastewater by
chitosan. J Colloid Int Sci 2005;286:3642.
[110] Hu ZG, Zhang J, Chan WL, Szeto YS. The adsorption of
acid dye onto chitosan nanoparticles. Polymer 2006;47:
58317.
[111] Hu ZG, Chan WL, Szeto YS. Dyeing silk with chitosan
nanoparticles. ATA J 2006;17:3840.
[112] Uzun I. Kinetics of the adsorption of reactive dyes by
chitosan. Dyes Pigm 2006;70:7683.
[113] Uzun I, Guzel F. Rate studies on the adsorption of some
dyestuffs and p-nitrophenol by chitosan and monocarboxymethylated(MCM)-chitosan from aqueous solution. Dyes
Pigm 2005;118:14154.
[114] Uzun I, Guzel F. Kinetics and thermodynamics of the
adsorption of some dyestuffs and p-nitrophenol by chitosan
and MCM-chitosan from aqueous solution. J Colloid Int
Sci 2004;274:398412.
[115] Uzun I, Guzel F. External mass transfer studies during the
adsorptions of some dyestuffs and p-nitrophenol onto chitosan
from aqueous solution. Turk J Chem 2004;28:73140.

ARTICLE IN PRESS
G. Crini, P.-M. Badot / Prog. Polym. Sci. 33 (2008) 399447
[116] Wen YZ, Liu WQ, Fang ZH, Liu WP. Effects
of adsorption interferents on removal of reactive red 195
dye in wastewater chitosan. J Environ Sci China 2005;17:
7669.
[117] Kofuji K, Qian CJ, Nishimura M, Sugiyama I, Murata Y,
Kawashima S. Relationship between physicochemical
characteristics and functional properties of chitosan. Eur
Polym J 2005;41:278491.
[118] Chang YC, Chang SW, Chen DH. Magnetic chitosan
nanoparticles: studies on chitosan binding and adsorption
of Co(II) ions. React Funct Polym 2006;66:33541.
[119] Bodnar M, Hartmann JF, Borbely J. Preparation and
characterization of chitosan-based nanoparticles. Biomacromolecules 2005;6:25217.
[120] Wan Ngah WS, Fatinathan S. Chitosan akes and
chitosan-GLA beads for adsorption of p-nitrophenol in
aqueous solution. Colloid Surf A Physicochem Eng Aspects
2006;277:21422.
[121] Knorr D. Dye binding properties of chitin and chitosan.
J Food Sci 1983;48:3641.
[122] Tien C. Adsorption calculations and modeling. Boston,
Newton, MA: Butterworth-Heinemann; 1994. 244pp.
[123] Langmuir I. The constitution and fundamental properties
of solids and liquids. JACS 1916;38:222195.
[124] Freundlich HMF. Uber die adsorption in losungen. Z Phys
Chem 1906;57:385471.
[125] Giles CH, Smith D, Huitson A. A general treatment and
classication of the solute adsorption isotherm. I. Theoretical. J Colloid Int Sci 1974;47:75565.
[126] Giles CH, DSilva AP, Easton IA. A general treatment and
classication of the solute adsorption isotherm. II. Experimental interpretation. J Colloid Int Sci 1974;47:76678.

447

[127] Lagergren S. Zur theorie der sogenannten adsorption


geloster stoffe. Kungliga Svenska Vetenskapsakademiens.
Handlingar 1898;24:139.
[128] Ho YS, McKay G. Kinetic models for the adsorption of
dye from aqueous solution by wood. Trans Chem Eng.
1998;76:18391.
[129] Weber WJ, Morris JC. Kinetics of adsorption on carbon
solution. J Sanitary Eng Div Am Soc Civ Eng 1963;89:
3159.
[130] Crank J. The mathematics of diffusion. 2nd ed. Oxford:
Clarendon Press; 1975. 414pp.
[131] Findon A, McKay G, Blair HS. Transport studies for the
adsorption of copper ions by chitosan. J Environ Sci
Health 1993;A28:17385.
[132] McKay G, Blair HS, Findon A. Adsorption of metal ions
by chitosan. In: Eccles H, Hunt S, editors. Immobilisation
of ions by bio-adsorption. Chichester, UK: Ellis Horwood;
1986. p. 5969.
[133] Ho YS. Review of second-order models for adsorption
systems. J Hazard Mater 2006;B136:6819.
[134] Bernardin Jr, FE. Experimental design and testing of
adsorption and adsorbates. In: Slejko FL, editor. Adsorption technology: a step-by-step approach to process
evaluation and application. New York, 1985. p. 3790
[chapter 2].
[135] Harry S. The theory of coloration of textiles. In: Johnson
A, editor. Thermodynamics of dye adsorption. 2nd ed.
West Yorkshire, UK: Society of Dyers and Colorists; 1989.
p. 255.
[136] Bailey SE, Olin TJ, Bricka M, Adrian DD. A review of
potentially low-cost adsorbents for heavy metals. Water
Res 1999;33:246979.

Вам также может понравиться