7
Colloids and Surfaces A: Physicoch em. Eng. Aspects 299 (2007) 146–15 2 Adsorptiv e removal of congo red, a carcinogenic textile dye by chitosan hydrobeads: Binding mechanism, equilibrium and kinetics Sandipan Chatterjee, Sudipta Chatterjee, Bishnu P . Chatterjee, Arun K. Guha  Department of Biological Chemistry, Indian Association for the Cultivation of Science, Jadavpur, K olkata 700032, India Received 21 November 2005; received in revised form 10 November 2006; accepted 15 November 2006 Available online 25 November 2006 Abstract Physico-chemic al investigation on adsorption of congo red, an anionic azo dye by chitosan hydrobeads has been carried out. Adsorption process has been found to be dependant on temperature with optimum activity at 30 C. Both ionic interaction as well as physical forces is responsible for binding of congo red with chitosan. Theoretical correlation of the experimental equilibrium adsorption data for congo red–chitosan hydrobeads sys temwoul dbe bes t explained by linea riz ed for mof La ngmuir iso the rm model.The kinetic res ult sfollowpseudosecond-orde r rat e equ ati on.pH of the experimental solution inuenced congo red adsorption inversely , and 20.0% of the dye could be desorbed from the loaded beads by changing thepH of thesolutiontoalkalinerange(pH 12.0). Both sodiu m chlor ideandsodiu m dodec yl sulf ate signi cant ly inue ncedthe adsor ptionprocess. © 2006 Elsevier B.V. All rights reserved. Keywords: Chitosan hydrobeads; Congo red; Adsorption; Binding mechanism; Zeta potential 1. Intro ducti on Annual production of textile dyes is estimated to be over 8 × 10 5 tonnes of which 10% are discharged as efuents [1]. Release of these dyes in water stream is aesthetically unde- sirable and has serious environmental impact. Due to intense colou r they reduc e sunli ght trans miss ion into wate r hence affe ct- ing aquatic plants, which ultimately disturb aquatic ecosystem [2]; in addition they are toxic to humans also. Due to the strong environmental regulation, removal of dyes is necessary before discharging into water bodies. Most of the textile dyes belong to the azo groups. Being electron decient in nature, they are less susceptible to oxidative catabolism [3] and generally non- biodegradable. Only a few can be degraded microbiologically under anaerobic condition [4] but in most cases with the pro- duction of carcinogenic amines [5] and mutagens [6]. Conventional procedures to remove dye from wastewater such as acti vat ed sludge, bacterial and chemi cal oxida tion, ozonisation or physical methods like membrane ltration, ion Corre spon dingauthor at: Depar tmen t of Biol ogic al Chemistry , Indian Asso - cia tio n forthe Cul tiv ati onof Sci ence, 2 A&BRajaS. C. Mulli ckRoad,Jada vpu r, Kolkata 700032, India. Tel.: +91 33 2473 5904/4971x502; fax: +91 33 2473 2805.  E-mail address: [email protected] (A.K. Guha). exchange or electrochemical techniques are either expensive or inadequate [7–13]. Solid phase extraction, commonly known as sorption, is the recent technology to remove dye from wastew- ater. Activ ated carbon is widely used as adsorbent [14] f or this purpose but high procuring and processing cost restricted its use in many countries. This promoted search for alternative cost effective adsorbent. Recently different low cost adsorbents including some industrial and agricultural wastes [15–22], such as y ash, fuller’s earth, waste red mud, bentonite clay, metal hydroxide sludge, peat, pith, cotton waste, rice husk, teakwood bark, etc. have been used but their effectiveness is limited and inferior to that of activated carbon. Adsorption capacity of the biological materials like coir pith [23], banana and orange peel [24] particularly towards anionic dye is very poor. Biological mat erials gen era lly contain hig h amo unt of cel lul ose and adsorb anion ic dye in a rev ersible proce ss in volv ing physi cal adso rption as against basic dyes where adsorption occurs there coulombic attraction and ion exchange process [25]. The biopolymer chitosan has gained importance in environ- mental biotechnology due to its very good adsorption capacity towa rds dye s and met al ion s [26–28]. More over,chitosan can be obtained at industrial scale by chemical deacetylation of crus- tacean chitin, bye product of seafood industry. Further, chitosan in the form of swollen beads has been found to exhibit better adsorption capacity than akes [29]. 0927-7757/$ – see front matter © 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.colsurfa.2006.11.036

Adsorptive of Congo

Embed Size (px)

Citation preview

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 1/7

Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152

Adsorptive removal of congo red, a carcinogenic textile dye by chitosanhydrobeads: Binding mechanism, equilibrium and kinetics

Sandipan Chatterjee, Sudipta Chatterjee, Bishnu P. Chatterjee, Arun K. Guha ∗

 Department of Biological Chemistry, Indian Association for the Cultivation of Science, Jadavpur, Kolkata 700032, India

Received 21 November 2005; received in revised form 10 November 2006; accepted 15 November 2006

Available online 25 November 2006

Abstract

Physico-chemical investigation on adsorption of congo red, an anionic azo dye by chitosan hydrobeads has been carried out. Adsorption processhas been found to be dependant on temperature with optimum activity at 30 ◦C. Both ionic interaction as well as physical forces is responsible for

binding of congo red with chitosan. Theoretical correlation of the experimental equilibrium adsorption data for congo red–chitosan hydrobeads

systemwould be best explained by linearized form of Langmuir isothermmodel.The kinetic results followpseudo second-order rate equation.pH of 

the experimental solution influenced congo red adsorption inversely, and ∼20.0% of the dye could be desorbed from the loaded beads by changing

thepH of thesolutionto alkaline range (∼pH 12.0). Bothsodium chlorideandsodium dodecyl sulfatesignificantly influencedthe adsorptionprocess.

© 2006 Elsevier B.V. All rights reserved.

Keywords: Chitosan hydrobeads; Congo red; Adsorption; Binding mechanism; Zeta potential

1. Introduction

Annual production of textile dyes is estimated to be over8× 105 tonnes of which 10% are discharged as effluents [1].

Release of these dyes in water stream is aesthetically unde-

sirable and has serious environmental impact. Due to intense

colour theyreduce sunlight transmission intowater henceaffect-

ing aquatic plants, which ultimately disturb aquatic ecosystem

[2]; in addition they are toxic to humans also. Due to the strong

environmental regulation, removal of dyes is necessary before

discharging into water bodies. Most of the textile dyes belong

to the azo groups. Being electron deficient in nature, they are

less susceptible to oxidative catabolism [3] and generally non-

biodegradable. Only a few can be degraded microbiologically

under anaerobic condition [4] but in most cases with the pro-

duction of carcinogenic amines [5] and mutagens [6].

Conventional procedures to remove dye from wastewater

such as activated sludge, bacterial and chemical oxidation,

ozonisation or physical methods like membrane filtration, ion

∗ Correspondingauthor at: Department of Biological Chemistry, Indian Asso-

ciation forthe Cultivationof Science, 2 A&BRajaS. C. MullickRoad,Jadavpur,

Kolkata 700032, India. Tel.: +91 33 2473 5904/4971x502;

fax: +91 33 2473 2805.

 E-mail address: [email protected](A.K. Guha).

exchange or electrochemical techniques are either expensive or

inadequate [7–13]. Solid phase extraction, commonly known as

sorption, is the recent technology to remove dye from wastew-ater. Activated carbon is widely used as adsorbent [14] f or this

purpose but high procuring and processing cost restricted its

use in many countries. This promoted search for alternative

cost effective adsorbent. Recently different low cost adsorbents

including some industrial and agricultural wastes [15–22], such

as fly ash, fuller’s earth, waste red mud, bentonite clay, metal

hydroxide sludge, peat, pith, cotton waste, rice husk, teakwood

bark, etc. have been used but their effectiveness is limited and

inferior to that of activated carbon. Adsorption capacity of the

biological materials like coir pith [23], banana and orange peel

[24] particularly towards anionic dye is very poor. Biological

materials generally contain high amount of celluloseand adsorb

anionic dye in a reversible process involving physical adsorption

as against basic dyes where adsorption occurs there coulombic

attraction and ion exchange process [25].

The biopolymer chitosan has gained importance in environ-

mental biotechnology due to its very good adsorption capacity

towardsdyes andmetal ions [26–28]. Moreover, chitosan can be

obtained at industrial scale by chemical deacetylation of crus-

tacean chitin, bye product of seafood industry. Further, chitosan

in the form of swollen beads has been found to exhibit better

adsorption capacity than flakes [29].

0927-7757/$ – see front matter © 2006 Elsevier B.V. All rights reserved.

doi:10.1016/j.colsurfa.2006.11.036

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 2/7

S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152 147

Objective of the present investigation is to study the efficacy

of chitosan in the form of hydrobeads to remove congo red,

a member of anionic azo dyes, from its aqueous solution and

to investigate the dye-chitosan interaction during adsorption in

single component system as well as in presence of different

industrial contaminants.

2. Materials and methods

2.1. Chemicals

Chitosan used in the present investigation was prepared from

shrimp shell by modified Hackmann procedure as describedear-

lier [30]. Average molecular weight and degree of deacetylation

of this chitosan were 398kDa and90.8%, respectively. All other

chemicals were of analytical grade and procured from BDH,

England.

2.2. Preparation of chitosan beads

Chitosan hydrobeads were prepared by drop-wise addition

of 3% chitosan solution (w/v) in 7% AcOH (w/v) to an alka-

line coagulating mixture (H2O: MeOH:NaOH:: 4:5:1, w/w) as

described by Mitani et al. [31]. The beads were spherical in

shape with an average diameter ∼2.5 mm as determined with

electronicmicrovernier caliper. Before use, chitosan beads were

kept for 30 min in water of pH adjusted to that required for

different experimental conditions.

2.3. Preparation of congo red solution and analytical

measurements

Stock solution of congo red (1 g/l) was prepared in deion-

ized water and desired concentrations of the dye were obtained

by diluting the same with water. Calibration curve of congo

red was prepared by measuring absorbance of different prede-

termined concentrations of the samples at λmax 497nm using

UV–vis spectrophotometer (Varian Carry 50). Concentration of 

dye in the experimental solution was calculated from the linear

regression equation(OD= 0.046× concentration of dyein mg/l)

of the calibration curve. Adsorption of dye by 1 g of chitosan in

the form of hydrobeads was determined by the following mass

balance equation.

q = (Ci−Cf )

×V 

1000×W 

where q represents theamountof dyeadsorbed,C i and C f are the

initial andfinal concentration (mg/ml) of dye after adsorption. V 

is the volume (ml) of experimental solution and W is the weight

(g) of chitosan in form of hydrobeads.

2.4. Scanning electron micrograph (SEM) studies

Chitosan beads beforeandafter adsorption of congo redwere

dried initially at 70 ◦C for 48 h and finally for 24 h after increas-

ing the temperature to 105 ◦C. Beads were kept in a desiccator

untiluse.The scanningelectron microscope (SEM)photographs

were taken in a Hitachi model S-415A scanning electron micro-

scope after coating the beads with gold by elctro-deposition

under vacuum.

2.5. Mesurement of zeta potential

Dry powder of chitosan beads (0.2 g) was added to 200 mldeionized water and pH of the suspension was adjusted with

0.1N NaOH or HCl to the desired level. The mixture was

stirred for 24h and then allowed to settle for 12h. Supernatant

containing some small particleswasused for zeta potentialmea-

surements in zetananoseries instrument(Malvern Instruments).

2.6. Equilibrium studies

Fifty millilitres aqueous solution of congo red (pH∼6.0, nat-

ural) with predetermined initial concentration (1–500 mg/l) was

taken in different 250-ml Erlenmeyer flasks. Swollen chitosan

hydrobeads (1 g) having water content of 96.4% were added to

each flask. The flasks were agitated (50 rpm) at 30± 1 ◦C for

48 h, which is expected to be far beyond the time required to

reach equilibrium. At the end of incubation, beads were sepa-

rated by filtration through glass wool and concentration of the

dye in the solution was estimated spectrophtometrically. The

same experiment was repeated at two different temperatures

(viz., 40 ◦C and 50 ◦C).

2.7. Adsorption dynamics

The above experiment was repeated with 500-mg/l congo

red solution at 30 ◦C and concentration of dye in the solution

was determined at different time intervals. Each data point wasobtained from individual flask and therefore no correction was

necessary due to withdrawal of sampling volume [32].

2.8. Influence of pH 

Influence of pH on the adsorption process was studied over a

range 3.0 to 12.0. Other conditions were the same as described

in kinetic study, and dye concentration was measured after 10 h.

2.9. Influence of sodium chloride, sodium dodecyl sulphate

and desorption studies

Influence of sodium chloride and sodium dodecyl sulphate

on dye adsorption was studied in similar way as described in pH

experiment. Congo red solution contained either sodium chlo-

ride (0.5M and 1 M) or sodium dodecyl sulfate (1% and 2%,

w/v); pH of the solution was maintained at 6.0.

After kinetic experiment, dye-loaded chitosan hydrobeads

(1 g) obtained by filtrationandwashing with water through glass

wool were taken in different 250-ml Erlenmeyer flasks. Fifty

millilitres water having pH adjusted to 8–12 or 50 ml ethanol,

methanol, propanol, ether, acetone was added to each flask. The

flasks were agitated (50 rpm) for 24h at30 ◦C and concentration

of the dye in the supernatant was estimated.

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 3/7

148 S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152

Fig. 1. SEM micrographs of Chitosan hydrobeads (a) congo red unadsorbed;

(b) congo red adsorbed.

2.10. Statistical analysis

Sigma Stat 3.0.1 statitistical analysis software was used to

analyze differentexperimental data.Energy minimizedstructure

and Connolly molecular area of the congo red molecule were

obtained using MOPAC and Chem Office 3D Ultra Software,

respectively.

3. Results and discussion

3.1. Scanning electron microscope study

Surface morphology of the dried chitosan beads (congo red

adsorbed and unabsorbed) was observed under scanning elec-

tron microscope at 380× magnification (Fig. 1). Beads weredried slowly in two steps, because quick drying may disturb

surface morphology of the beads due to fast evaporation. SEM

photographs reveal that thesurfaceof unadsorbed beadswas typ-

ically wrinkled polymeric network with irregular pores, which

after adsorption of the dye became slightly flat and pores were

not visible. This predicts that thin layer of the dye has covered

the entire external surface of the beads.

3.2. Equilibrium studies

Adsorption of congo red from its aqueous solution by chi-

tosan hydrobeads at different temperature is presented in Fig. 2.

Fig. 2. Variation of equilibrium uptake of congo redwith temperature at various

initial concentration of congo red. Results represent average of five replicate

experiments.

The results show that adsorption of the dye was maximum at

30 ◦C. Adsorption of a solute from solution phase onto solid liq-

uid interface occurs by dislodging the solvent molecule (water)

from interfacial region [33]. Probably with rise in temperature

from 20 ◦C to 30 ◦C, the interaction between solvent and solid

surface reduced exposing more number of adsorption sites as

reported earlier [34]. Thisenhanced the possibility of interaction

between dye and chitosan. In addition, at elevated temperature,

anincrease in free volume occurred,whichalsofavoured adsorp-

tion [35]. From 30 ◦C onwards, adsorption started to fall sharply

indicating exothermic nature of the process. Moreover, at high

temperature,the physical interaction between congoredandchi-tosan weakened due to weakening of hydrogen bonds and van

der Waals interaction. At 30 ◦C, ∼92.59mg of congo red was

removed by 1 g of chitosan in the form of hydrobeads at pH

6.0, which was much higher in comparison to other adsorbent

reported, such as waste wollestonite (1.21 mg/g) [36], biogas

waste slurry (9.5 mg/g) [37], waste banana pith (20.29 mg/g)

[38], paddy straw (1.01mg/g) [39], Fe (III)/Cr (III) (44.0mg/g)

[40], waste orange peal (22.44 mg/g) [20], waste red mud

(4.05mg/g) [41], activated carbon prepared from coir pith [23],

Neem leaf powder (41.2 mg/g) [42], Aspergillus niger biomass

(14.16 mg/g) [43]. Surprisingly, both granulated (13.80 mg/g)

and powdered activated carbon (16.81 mg/g) [43] have beenfound to be less efficient to chitosan. Biological materials gen-

erally contain negatively charged cellulose that repels anionic

dyes by strong electrostatic repulsion resulting in poor adsorp-

tion of dye [25]. This may be one of the reasons for their low

adsorption capacity towards congo red. On the other hand, chi-

tosan being polycationic in nature attracts congo red, an anionic

dye, and thus increased dye adsorption.

3.3. Adsorption isotherm

Adsorptionisothermmodeldescribes howadsorbate interacts

withadsorbents, and comprehensiveunderstanding of the nature

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 4/7

S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152 149

of interaction is essential for the most effective use of the adsor-

bent. It is thus essential to correlatethe equilibriumdata byeither

theoretical or empirical models for designing a perfect oper-

ating adsorption system for industrial effluents. In the present

investigation, the equilibrium data were analyzed according to

the linear form of Freundlich (Eq. (1)) and Langmuir (Eq. (2))

model of adsorption isotherm.

ln qe = ln KF +1

nln Ce (1)

Ce

qe=

1

KL+

aL

KLCe (2)

where C e is the remaining dye concentration (in mg/l) in solu-

tion, qe the dye concentration (in mg/g) in the adsorbent, K F the

Freundlich constant and 1/ n is the heterogeneity factor. While

aL and K L are the Langmuir constants.

Freundlich isotherm describes the heterogeneous system and

reversible adsorption. This is not restricted to the monolayer

formation. This model predicts that dye concentration in theadsorbent will increase with increase in dyeconcentration in the

solution. Onthe other hand in the Langmuir model, it is assumed

that intermolecularforces decreaserapidly withdistanceandthis

leads to the prediction that coverage of adsorbent by adsorbet is

of monolayer type. Further, it is assumed that once a particular

site of theadsorbent is occupied by an adsorbet molecule,no fur-

ther adsorption takes place at that site. Theoretically, adsorbent

has finite number of sites and once all these sites are occu-

pied by adsorbet further adsorption cannot take place. Plotting

of experimental data into linearized form of Freunlich model

(Fig. 3a) shows deviation from linearity if the whole concen-

tration range is taken into consideration. However, if the wholeconcentration range is divided into three linear regions,good fits

to the experimental data are noted. This is probably (i) due to

the presence of different groups such as amino, hydroxyl, acetyl

in the chitosan molecule producing irregular energy distribution

in the adsorbent (chitosan bead) surface or (ii) by a purely phys-

ical adsorption [44]. However, Langmuir model correlated well

(correlation coefficient ∼1.0) when only one line was used to

represent the entire range of experimental data (Fig. 3b). The

general shape of the isotherm curve including sharp curvature

near to the saturation point and short equilibrium time are also

very characteristics of a Langmuir equilibrium with high sorp-

tion capacity. In the Langmuir model, slope of the curve K L / aL

(≡Qo) represents the theoretical monolayer saturation capacityof the adsorbet on the adsorbent and aL is related to the adsorp-

tion energy. High value of aL at 30 ◦C obtained in the present

case indicates strong binding and with rise in temperature; the

value of aL reduced significantly indicating the binding force

weakenedat higher temperature. Theexperimental findings also

corroborated the saturation capacities predicted by Langmuir

equation at different temperatures.

Essential features of theLangmuir isotherm canbe expressed

in terms of dimensionless equilibrium parameter RL, which is

1/(1+ aLC 0) where aL is the Langmuir constant as described

above and C 0 is the initial dye concentration. Value of  RL

less than one but greater than zero indicates favourable adsorp-

Fig. 3. (a) Freundlich plot of congo red–chitosan beads equilibrium exper-

imental data. (b) Langmuir’s plot of congo red–chitosan beads equilibrium

experimental data. R represents the correlation coefficient and the equation

represents the regression equation of the linear portion of the curve.

tion. RL values calculated at different initial dye concentrations

are well within the defined range and this indicates favourable

adsorption. All the RL values are very close to lower accept-

able range (Table 1) indicating high degree of irreversibility of 

the process. Constants related to the Freundlich and Langmuir

model are presented in Table 2.

3.4. Adsorption dynamics

Fig. 4 shows that adsorption of congo red on chitosan beads

takes place at relatively fast rate in the first 10 min and then

Table 1

 RL value for the adsorption of congo red onto chitosan hydrobeads

Temperature (◦C) Initial concentration of congo red in solution (mg/l)

50 100 200 300

30 0.071 0.037 0.019 0.013

40 0.189 0.104 0.055 0.037

50 0.250 0.143 0.077 0.053

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 5/7

150 S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152

Table 2

Constants of Freundlich’s and Langmuir’s equation

Temperature (◦C) Constant Region of the Freundlich’s curve

Lower Middle Higher

Freundlich’s equation

20K F 7.79 11.39 66.69

1/ n 0.734 0.75 0.06

30K F 10.26 16.64 72.18

1/ n 0.600 0.703 0.045

40K F 5.13 8.15 66.27

1/ n 0.732 1.23 0.903

50K F 3.75 4.42 58.56

1/ n 0.832 0.873 0.048

Temperature (◦C) aL (mg−1) k L (g−1) k L /a L ( Qo, mg/g)

Langmuir’s equation

20 0.101 9.47 93.40

30 0.257 24.07 93.71

40 0.091 7.99 87.8

50 0.063 5.10 80.76

gradually decreased to reach equilibrium around 7 h. Initial fast

rate isprobably dueto greater availability of binding sitesaround

the surface of the chitosan hydrobeads.

Linear form of pseudo first-order and second-order kinetic

rate models, were studied with the kinetic experimental data

to examine the controlling mechanism of adsorption. Table 3

shows that correlation coefficient for both forms of rate models

are almost same but calculated and experimental equilibrium

uptake as well as linearity of the experimental data fit well to

pseudo second-order model. This has further been confirmed by

experimental and generated kinetic rate profiles (Fig. 4).

Chitosan bead is an open porous system; as such there is

every possibility that thedyemaybe transported within thebead

through its porous network. However, cross-section of congo

red-adsorbed beads showed no sign of intraparticle diffusion.

Molecular shape and size are the prime parameters in the intra-

Fig. 4. Effect of contact time on the congo red–chitosan equilibrium at different

initial dye concentration.

Table 3

Constants of different rate models

Experimental

C o (mg/l) 200.0

C e (mg/l) 133.34

qe (exp) (mg/g) 92.59

First-order

 R 0.974k 1 7.5× 10−3

qe (cal) (mg/g) 65.77

Second-order

 R 0.998

k 2 1.64×10−4

qe (cal) (mg/g) 99.0

particle diffusion process. The software package CS Chem 3D

ultra was used to calculate the Connolly molecular surface area

of congo red molecule, which gives a fairly accurate prediction

of the area occupied by spatial orientation of various functional

groups and structures of a molecule. Large molecular surfacearea (557.6 ´̊A2) predicts large dimension of the molecule and

hence unable to diffuse through the porous network of the beads

used in the present investigation. This implies that intraparticle

diffusion has no role in the mass transfer of this process.

3.5. Effect of pH and adsorption mechanism

Adsorption of congo red by chitosan had been found to be

dependant on pH of the solution (Fig. 5), which increased with

decrease in pH. In order to understand the role of pH on the

adsorption process, the surface charge of chitosan beads was

determined by measuring zeta potential. Surface charge of chi-tosan is positive in acidic pH, which decreases gradually with

increase in pH and pass through zero potential at pH 6.4.

In acidic pH range, surface charge of the adsorbent increases

mainly due to increased protonation of amine group (–NH3+)

of chitosan. Congo red is an acidic dye and contains negatively

charged sulfonated group (–SO3− Na+). Higher adsorption of 

Fig. 5. Effect of pH on congo red adsorption by chitosan hydrobeads.

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 6/7

S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152 151

Fig. 6. Cogo red–chitosan interaction: (i) ionic interaction (involves when pH of experimental solution is below 6.4), (ii) hydrogen bonding between hydroxyl group

of chitosan and electronegative residues in the dye molecule, and (iii) Yosida H- bonding between hydroxyl group of chitosan and aromatic residue in dye.

the dye at lower pH is probably due to the increase in elec-

trostatic attraction between negatively charged dye moleculeand positively charged amine group of chitosan. However, at

pH 6.4 where surface charge of chitosan hydrobeads is neu-

tral, adsorption of the dye in this pH range may be attributed to

physical forces only. The three-dimensional energy minimized

structure of congo red as drawn with Chem Office 3D ultra soft-

ware appears to be flat. Thus attachment of congo red molecule

on the surface of chitosan beads is of flat layer type. Hence there

is every possibility of hydrogen bond formation between some

of the molecular components of congo red such as N, S, O, ben-

zene ring and CH2OH groups of the chitosan molecule [45].

At pH above 6.4, surface charge of chitosan beads is negative

which hinders the adsorption by electrostatic force of repul-sionbetween thenegatively chargeddye moleculeand adsorbent

(chitosan hydrobeads); but appreciable amount of adsorption in

this pH range suggests strong involvement of physical forces

such as hydrogen bonding, van der Waals force, etc. in the

adsorption process. Fig. 6 represents all possible interactions

between chitosan and congo red.

3.6. Effect of additives on adsorption and pH on desorption

 phenomena

Sodiumsalt is oftenused as stimulator in dying industries and

sodium dodecyl sulphate, an anionic surfactant is often presentin the effluent of dye industries; hence their effect on adsorp-

tion process was studied. Sodium chloride at the concentration

of 0.5 M and 1 M reduced the dye adsorption by 14.4% and

28.0%, respectively(Fig.7). Sodiumchlorideis knownto reduce

ionic interaction [46] and thus supports our prediction regard-

ing themechanismof congo redadsorption particularly in acidic

pH on chitosan. Sodium dodecyl sulphate like sodium chloride

inhibited adsorption of congo red (Fig. 7).

Investigation on the desorption of the dye from the adsorbent

is necessary for its reuse and also to understand the mechanism

of adsorption. Table 4 shows that amount of dye desorbed from

the loaded chitosan beads increased with increase in pH of the

Fig. 7. Effect of sodium chloride and sodium dodecyl sulphate over the adsorp-

tion of congo red by chitosan hydrobeads.

Table 4

Desorption of congo red from dye-loaded beads

pH % of dye desorbed

8.0 15.4

10.0 18.0

11.0 19.0

12.0 21.2

eluent. Low desorption of dye with pH change indicates thatphysical forces such as hydrogen bonding and van der Waals

force play the predominate role in the removal of congo red by

the chitosan hydrobeads because the change in pH only affects

the surface charge of the adsorbent. Organic solvents, except

acetone were ineffective in desorption of congo red from loaded

beads. However, acetone desorbed the dye but the extent was

only 25%.

4. Conclusion

Following conclusions may be drawn from the present inves-

tigation.

8/8/2019 Adsorptive of Congo

http://slidepdf.com/reader/full/adsorptive-of-congo 7/7

152 S. Chatterjee et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 299 (2007) 146–152

1. Chitosan beads is a good adsorbent for the removal of congo

red from its aqueous solution and 1g of chitosan in the form

of hydrobeads can remove ∼93 mg of the dye at natural pH

(∼6.0).

2. Langmuir adsorption isotherm model fitswell to describe the

equilibrium adsorption isotherm data.

3. Adsorption process is dependant on pH and temperature and

followssecond-orderkinetics rate model todescribe themass

transfer mechanism.

4. Sodium chloride and sodium dodecyl sulphate (SDS) inhibit

the adsorption process.

5. Physical forces as well as ionic interaction are responsible

for binding of congo red with chitosan.

Acknowledgements

Authors are thankful to Dr. P.C Banerjee of Indian Institute

of Chemical Biology, Kolkata for valuable suggestions. Sandi-

pan Chatterjee wishes to acknowledge Council of Scientific and

Industrial Research (CSIR), New Delhi, for providing fellow-ship.

Appendix A. Supplementary data

Supplementary data associated with this article canbe found,

in the online version, at doi:10.1016/j.colsurfa.2006.11.036.

References

[1] H. Zollinger, Colour Chemistry-Synthesis, Properties and Application of 

Organic Dyes And Pigments, VCH Publishers, New York, 1987.

[2] S.A. Figueiredo, R.A. Boaventura, J.M. Loureoro, Sep. Purif. Technol. 20

(2000) 129.[3] C.A. Fewson, Trends Biotechnol. 6 (1988) 148.

[4] S. Chinwetkitvanich, M. Tuntoolvest, T. Panswad, Water Res. 34 (2000)

2223.

[5] U. Mayer,Biodegradation of SyntheticOrganicColorant FEMSSymp, vol.

12, 1981, p. 371.

[6] K.T. Chung, C.E. Cerniglia, Mutat. Res. 277 (1992) 201.

[7] P.C. Vandevivere, R. Bianchi, W. Verstraete, J. Chem. Tech. Biotechnol. 72

(1998) 289.

[8] G.M. Walker, L.R. Weatherley, Water Res. 31 (1997) 2093.

[9] R. Ganesh, G.D. Boardman, D. Michelsen, Water Res. 38 (1994) 1367.

[10] W. Chu, S.M. Tsui, Chemosphere 39 (1999) 1667.

[11] Y.M. Slokar, A.M. Le, Meyer, Dyes Pigm. 37 (1997) 335.

[12] G. Mishra, M.A. Tripathy, Colourage 40 (1993) 35.

[13] U.B. Ogutveren, S. Kaparal, J. Environ. Sci. Health A 29 (1994) 1.

[14] N.M. Nasser, M.J. El-Geundi, J. Chem. Technol. Biotechnol. 50 (1991)

257.

[15] G. Mc Kay, M.S. Otterburn, J.A. Aga, Water Air Soil Pollut. 24 (1985)

307.

[16] C. Namasivayam, K. Ranganathan, Res. Ind. 37 (1992) 165.[17] S. Netpradit, P. Thiravetyan, S. Towprayoon, Water Res. 37 (2003) 763.

[18] G. Mckay, S.J. Allen, Can. J. Chem. Eng. 58 (1980) 521.

[19] C. Namsivayam, N. Kanchana, Chemosphere 25 (1992) 1691.

[20] C. Namasivayam, N. Muniasamy, K. Gayathri, M. Rani, K. Ranganathan,

Biores. Technol. 57 (1996) 37.

[21] G. Mckay, M. El-Geundi, N.M. Nassar, Water Res. 21 (1987) 1513.

[22] G. McKay, G. Ramprasad, P. Mowli, Water Air Soil Pollut. 29 (1986)

273.

[23] C. Namasivayam, D. Kavitha, Dyes Pigm. 54 (2002) 47.

[24] G. Annadurai, R.L. Juang, D.J. Lee, J. Hazard. Mater. B92 (2002) 263.

[25] A. Bousher, X. Shen, R.G.J. Edyvean, Water Res. 31 (1997) 2084.

[26] H. Yoshida, T. Takemori, Water Sci. Technol. 35 (1997) 29.

[27] L. Jin, R. Bai, Langmuir 18 (2002) 9765.

[28] E. Guibal, Sep. Purif. Technol. 38 (2004) 43.

[29] H.K. No, S.P. Meyers, Env. Cont. Toxico. 163 (2000) 1.[30] S. Chatterjee,A.K. Guha, B.P. Chatterjee, Ind.J. Chem. Technol. 10 (2003)

350.

[31] T. Mitani, N. Fukomuro, C. Yoshimoto, H. Ishii, Agric. Biol. Chem. 55

(1991) 2419.

[32] S. Chatterjee, S. Chatterjee, B.P. Chatterjee, A.R. Das, A.K. Guha, J. Col-

loid Interf. Sci. 288 (2005) 30.

[33] D.K. Chattoraj, T. Imae’, A. Mitra, Langmuir 20 (2004) 4903.

[34] T.N.decastro Dantas’,A.A. Dantas Neto, A.M.C.P. de,E.L. Moura, Barros,

de. Paiva Telemeco, Langmuir 17 (2001) 4256.

[35] Y.S. Ho, G. McKay, Chem. Eng. J. 70 (1998) 115.

[36] V.N. Singh, G. Mishra, K.K. Panday, Indian J. Technol. 22 (1984) 70.

[37] C. Namasivayam, R.T. Yamuna, J. Chem. Technol. Biotechnol. 53 (1992)

153.

[38] C. Namasivayam, M. Kanchana, Perntiak J. Sci. Technol. 1 (1993) 32.

[39] N. Deo, M. Ali, Ind. J. Env. Prot. 13 (1993) 496.

[40] C. Namasivayam, R. Jeyakumar, R.T. Yamuna, Waste Manage. 14 (1994)

643.

[41] C. Namasivayam, D.J.S.E. Arasi, Chemosphere 34 (1997) 401.

[42] K.G. Bhattachryya, A. Sharma, J. Env. Mannage. 71 (2004) 217.

[43] Y. Fu, T. Viraraghavan, Adv. Env. Res. 7 (2002) 239.

[44] W. Fritz, E.U. Schlunder, Chem. Eng. Sci. 36 (1981) 721.

[45] Y.C. Wong, Y.S. Szeto, W.H. Cheung, G. McKay, Langmuir 19 (2003)

7888.

[46] M.S. Chiou, H.Y. Li, Chemosphere 50 (2003) 1095.