You are on page 1of 83

CHAPTER 4

Chapter 4

Kinetic and Equilibrium Models for the Sorption


of Ueavy Metals from Aqueous Solutions by
Polymer-Grafted Banana Stalk

4.1 Introduction

The ability of inexpensive agricultural residues and forestry waste products as


adsorbents to adsorb heavy metal ions has received considerable attention for the
development of an efficient, clean and cheap technology for wastewater treatment
(Sharma and Forster, 1994; Wafwoyo et al., 1999; Petemala et al., 1999; Ho et al.,
2001; Reddad et al., 2002; Montanher et al., 2005; Nasemejad et al., 2005). These
materials possess some unique properties caused by oxygen-containing functional
groups on their surfaces. Unlike synthetic ion exchange resins these materials are
cheap and do not require complex regeneration and conservation. A problem with
the use of these materials for wastewater treatment is that they release phenolics,
which contaminate the treated water itself leading to pollution and reduction in
adsorption potential of the materials. Efforts have been made to prevent the
leaching of phenolic compounds from the adsorbent by numerous chemical
treatments (Okieimen et al., 1991; Simkovic and Laszlo, 1997; Wartelle and
Marshall, 2001). Chemical modifications of agricultural residues via esterification,
cross linking and grafting have been extensively studied. It has been reported that
the introduction of reactive functional groups in the backbone of cross linked
lignocellulosic substrates resulted in products that are capable of removing heavy
metals from industrial wastewaters (Khalil et al., 1991). New research in this
direction indicates that graft copolymerization on adsorbent materials followed by
98
Kinetic and Equilibrium Models ...........

functionalization prevents the leaching of organic compounds, such as lignin,


pectin, tannin etc. into the solution and also improving their binding properties
(Raji and Anirudhan, 1998). The advantage of this method also lies on the increase
of stability of the adsorbent materials, which is an important aspect of the
commercial development of biosorbent materials.
The impending portions of this chapter deal with the methods and optimum
conditions adopted for the synthesis of a new adsorbent, polyacrylamide-grafted
banana stalk bearing carboxylate functional group (PGBS-COOH). This work also
includes a detailed study of the characterization of the newly developed adsorbent,
and the prospects of employing PGBS-COOH for the removal of Pb(II), Co(II),
Hg(II) and Cd(II) ions from water. A probable mechanism for graft
copolymerization and surface functionalization is also suggested. This
investigation also focuses on the metal removing capacity, and the various factors
affecting the adsorption process at kinetic and equilibrium conditions. The
equilibrium isotherm data were analyzed using three isotherm models, Langmuir,
Freundlich and Scatchard, to determine the best fit equation for the sorption of
metals on the PGBS-COOH. A comparative study with a commercial cation
exchanger, Ceralite IRC-50, having carboxylate functional group is conducted with
a view to highlight the superiority of the newly synthesised adsorbent. Real,
synthetic or simulated wastewaters of the different metals were also treated by the
adsorbent to demonstrate its efficiency in removing the metals in the presence of
other inorganic contaminants. Desorption studies were performed for the recovery
of heavy metals and to check the utility of spent adsorbents.

4.2 Polymer Grafting


Graft polymerization of acrylamide (AAm) onto BS was conveniently carried out
in water using ferrous ammonium sulphate/H202 initiator system (Huang et al.,
1992). The polymerization was started by adding monomer (AAm). The optimum
99
Chapter4

conditions for graft copolymerization of acrylamide onto BS were investigated by


studying the effect of time, initiator, monomer and temperature on the percentage
graft yield. The graft yield (%) was calculated as follows

Wt of copolymer - Wt of substrate
Graft yield (%) = x 100 (4.1)
Wt of substrate

4.2.1 Effect of temperature

The graft copolymerization of acrylamide (AAm) onto Banana stalk (BS) was
carried out at different temperatures ranging from 30 to 70 °c (Figure 6). The graft
yield increases with increase of temperature. This may be attributed to the
following factors: (i) enhancement of swellability and thereby increasing the
surface area (iiJ increase in the mobility of the monomer and the initiator resulting
in the vicinity of the BS particles (iii) higher rate of diffusion of monomer and
initiator from the solution phase to the solid phase (iv) it is most probable that the
accompanied increase in the number of free radicals on raising the temperature
upto 70 °c favours initiation rather than termination of polymerization. The results
also show that graft yield increases at the beginning and subsequently levels off.

4.2.2 Effect of FeAmS04

The percentage graft yield increases with increase in the concentration of


FeAmS04 up to 8.0 x I 0-4 M in the case of higher temperatures (above 40 uC) and
thereafter graft yield decreases slightly (Figure 7). In the case of 30 °c, percentage
graft yield increases up to 8.0 x 10-4 M and then assumes steady values. The
possible reason of the above trend is the formation of a greater number of active
sites by the increase in FeAmS04 concentration up to a limiting value. Beyond this,
concentration of the initiating free radical may also decrease due to the mutual
100
Kinetic and Equilibrium Models ...........

combination. Therefore 8.0 x 104 M FeAmS04 concentration has been used for the
study of all other effects.

..
'\40
& : 30 °C
120 • : so 0c

-

100
e : 60 °C

-�

.··.:s!� 80

60
"'
. it:
FeAmS04 : 8.0x 10"4M
40 2
H202 :L2 X 10" M
AAm : 25 lefL
20
BS : 5 lefL

0
0 50 100 150 200 250 300 350
Time (min)

Figure 6
Effect of reaction time and temperature on the graft copolymerization
of AAm onto BS
101
Chapter4

160

•••
°
A :30 C

-
:50 °C
: 60 °C
120
:70 °C

.i 80

: 1.2 X 10-2 M
C,
40 : 25 g,'L
: 5 g,'L
Time : 4h

0
0 2 4 6 8 10 12 14
4
[FeAmS04) x 10 (M)

Figure 7
Effect of concentration of ferrous ammonium sulphate
on the graft coplymerization of AAm onto BS

Figure 8 shows that graft yield increases with increase of H202 up to 1.2 x 10-2 M
concentration and remains constant thereafter. This may be due to the enhanced
rate of termination, which balances the rate of propagation (Pradhan et al., 1982).

4.2.4 Effect ofAAm

The effect of monomer was studied by varying the concentration from 5.0 to 40.0
g/L (Figure 9). The graft yield increased with increase of the monomer
concentration. The enhancement of grafting with the increase of monomer
concentration could be attributed to the following: First, the complexation of BS
102
Kinetic and Equilibrium Models ...........

with monomer required for enhancing monomer activity would be favoured at


higher monomer concentration a second cause might be the gel effect, resulting
from an enhanced solubility of polyacrylamide in its own monomer (Moharana and
Tripathy, 1991 ). This would, consequently, increase the viscosity of the reaction
medium and a reduced rate of termination by the coupling of the growing polymer
chain.

••
120
:30 °C

- ••
100 :50 °C
: 60 °c
:70 °C
'$. 80

.� 60

: 8.0x 10"4M
c., 40 AAm :25 gL
BS : 5 gL
20 Time :4 h

0.0 0.5 1.0 1.5 2.0


2
(H202] X 10 (M)

Figure 8
Effect of concentration of H202 on the graft
copolymerization of AAm onto BS
103
Chapter4

...
160

•••
:30 °C
140 :50 °C
:60 °C
120 : 10 °c
100

-� 80

60
C,
FeAmS04 : 8.0x 10"4M
40
H202 : l.2x 10·2M
20 BS : 5 g'L
Time : 4h
0
0 10 20 30 40 50
[AAm] (g/L)

Figure 9
Effect of monomer concentration (AAm) on the
graft copolymerization of AAm onto BS

4.3 Mechanism of Graft Polymerization

The mechanism of graft polymerization of lignocellulose has been studied


extensively by earlier workers (Moharana and Tripathy, 1991; Sreedhar and
Anirudhan, 2000). Huang et al. (1992) reported that lignin is more sensitive to
polymerization than is cellulose. Lignins generally contain a relatively high
concentration of p-hydroxyphenylpropane group (Vazquez-Torres et al., 1993) as
well as variable quantities of syringyl and guaiacyl groups (Figure I 0). Though the
104
Kinetic and Equilibrium Models ...........

composition of lignocellulose and the structure of lignin are rather complicated, the
sensitive component for polymerization might be the phenolic hydroxyl groups of
lignin. A mechanism of graft polymerization of AAm onto lignocellulose in BS,
initiated by Fe2+/H202 redox system, is proposed as follows: the ferrous ions first
react with H202 to produce ferric ions and hydroxyl radicals. The hydroxyl radicals
might attack lignocellulose to produce macroradicals. The newly produced ferric
ions then react with the phenolic hydroxyl group of the lignin to produce phenolic
radicals and ferrous ions. Then grafting of AAm takes place resulting in the
formation of PGBS.

I I I
_f_?_? OH

i. R1,R2 = H
ii. R1 = H, R2 = OCH3
iii. R1, R2 = OCH3

Figure 10
Structure of the lignin backbones for (i) p-hydroxyphenylpropane
(ii) guaiacyl and (iii) syringyl group

Initiation
2+ 3+
2 2 +Fe � Fe +OH- +OH•
HO (4.2)

Fe 2 + +
105
Chapter4


-BS +AAm -BS-AAm
• (4.4)

where -BS- is the phenolate ion of the lignin, -Bs• is the phenolate radical,
andAAm is the monomer.

Propagation

-BS-AAm• +AAm -BS-(AAmh• (4.5)

-BS-(AAm)n• (4.6)
- BS-(AAmJn -1• + AAm

Termination

-BS-{AAmJi• +-BS-(AAmJi• ---) PGBS (4.7)


(Graftcopolymer)
2+
- BS-{AAmJi • + Fe3+ ---) Fe +-BS-(AAmJi (4.8)

4.4 Adsorbent Characterization

4.4.1 Infrared spectra

Figure 11 shows the FTIR spectra of BS and PGBS-COOH. The broad absorption
band at about 3400 cm- 1 for BS is attributable to the sum of the contribution from
hydrogen bonded 0-H stretching vibration from cellulose structure and hydroxyl
group from polyphenols originally present in BS. The Band observed at 2945 cm-1
was attributed to C-H stretching from CH2 groups. The characteristic bands at
676 cm- 1 for BS and 653 cm- 1 for PGBS-COOH arise from 13-glucosidic linkage.
The broad absorption band at 3300 cm- 1 observed in the IR spectrum of
106
Kinetic and Equilibrium Models ...........

PGBS-COOH representing the overlap of 0-H, C-H, N-H and C-0 stretching
vibrations (Nakamoto, 1988). The peaks at 1626 cm- 1 (amide carbonyl) and 1568
cm- 1 for PGBS-COOH were due to the presence of aliphatic amide group and the
weak band at 1180 cm- 1 was due to the N-H stretching vibration. The participation
of 0-H groups in the grafting of AAm on BS was suggested by the variation
observed in broad absorption band due to 0-H stretching from 3400 cm- 1 in BS to
3300 cm-1 in PGBS-COOH. The new peaks at 1696 cm- 1 (vc=o) and 1455 cm- 1 (vc-o)
indicate the existence of -COOH groups in PGBS-COOH. These observations
clearly indicate the formation of polymeric chain (back bone) in the PGBS-COOH.

3400 2400 1400 400


Wavenumbers (cm"1)

Figure 11
FTIR spectra of BS and PGBS-COOH
107
Chapter4

4.4.2 Thermal analysis

The thermal behaviour of BS and PGBS-COOH was examined by the study of


their primary TG and DTA analysis (Figure 12). It was observed that the thermal
stability of PGBS-COOH is higher than that of the original BS. In the case of
original BS, the decomposition temperature, T0 is approximately 200 °c and in the
case of PGBS-COOH, T0 is 280 °c. In general, the thermal stability of
PGBS-COOH is comparable to typical phenol-formaldehyde resins (Vazquez­
Torres et al., 1993). The weight loss of BS was relatively higher than that of
PGBS-COOH. The TG curve showed an initial small weight loss of 10.0 % for BS
starting at 70°c and ending up at 120 °c, whereas in the case of PGBS-COOH a
weight loss of 7.0 % starting at 90 °c and ending up at 120 °c due to the loss of
physically adsorbed water. BS exhibits an endotherm close to 100 °c. No other
features visible at higher temperatures. The temperature break observed for
PGBS-COOH in the range of 120-180 °c indicated the loss of water molecule
which is possibly associated (coordinated) with the carboxylate group on the
adsorbent surface. The DTA curve shows a broad endothermic peak in the range of
90-180 °c, coinciding with loss of both physically adsorbed and coordinated water
in the adsorbent system, and a major endothermic peak at 420 °c. The temperature
for 10 % weight loss (T1 0) and temperature of maximum rate of degradation (Trnax)
are the two main criteria used to indicate the thermal stability of polymers (Mathew
and Pillai, 1993). The higher the values of Tio and Tmax, the higher the thermal
stability of the system. The values of T 10 and Tmax for PGBS-COOH were found to
be 190 and 450 °c respectively, indicating the high thermal stability of the
adsorbent.
108
Kinetic and Equilibrium Models ...........

100 0.2
-:BS
80 - : PGBS-COOH 0
0.1
l
-
-
60 [>-

l!
.21)
0 --=�
(')
40

20
-0.1 l
0
0 -0.2
0 200 400 600

Temperature (C)

Figure 12
TG and OTA curves of BS and PGBS-COOH

4.4.3 X-ray diffraction studies

X-ray diffraction patterns of the BS and PGBS-COOH are shown in Figure 13. It is
obvious that there is no new phase formation due to the grafting of polyacrylamide
on BS. However, results suggest that BS and its grafted products have their
corresponding crystalline structure. Native BS shows scattering angles at 2 B = 13°,
27° and 34°. XRD of BS indicated that BS has a crystalline domain of cellulose
substrate. In the crystalline regions, cellulose molecules (major component of BS)
are arranged in ordered lattices, in which hydroxyl groups are bonded by strong
secondary forces. Diffraction maxima at 20° and 34° have been attributed to the
crystalline region of cellulose. XRD studies showed that on graft polymerization, a
significant decrease in crystallinity took place. The broad peak centered at 21 °
109
Chapter4

appears on PGBS-COOH due to various lattice planes and is a combination of


peaks of various intensities (lower ordered fractions) (Young, 1986). Thus some
rearrangement in the morphology of cellulosic chains in PGBS-COOH would
occur by grafting. The decrease in crystalline domains in PGBS-COOH resulted in
the loss of tensile strength of grafted chain and consequently enhanced the free
mobility of grafted chain. The change is speculated to take place by small lateral
shifts of the chain segments which are loosened from one another by grafting,
giving rise to an increase in the lower ordered fractions. The free mobility of
grafted chains was enhanced also due to the breakdown of hydrogen bonding
network structure during chemical reaction.

100
-:BS
80 - : PGBS-COOH

= 60

·=- 40

20

0
10 20 30 40 50
6
2-Theta-Scale ( )

Figure 13
X-ray diffraction patterns of BS and PGBS-COOH

4.4.4 Scanning electron micrographs

The surface characteristics of BS and PGBS-COOH were examined using SEM


(Figure 14). The surface morphology of the BS is different from that the
PGBS-COOH. The SEM of BS has unit cells running longitudinally with parallel
110
Kinetic and Equilibrium Models ...........

orientations. The intercellular gaps, in the form of longitudinal cavities, can be


clearly seen as the unit cells are partially exposed. In order to hold the unit cells
firmly in the stalk fibers the intercellular gaps are filled by the binder lignin and
fatty substances. The SEM picture of PGBS-COOH clearly shows that
polyacrylamide (PAAm) grafts were deposited on the surface of the unit cell and
this caused the surface to become more or less smooth as compared to untreated
BS. PGBS-COOH was seen to be a homogeneous continuous matrix and there was
no evidence for the existence of separate phases. The overall fracture mode was
ductile and clearly corresponded to the PAAm grafts phase (Sawyer and Grubb,
1987). The deposits of PAAm grafts on the surface of the unit cell resulted in deep
voids in between the overgrowths of the unit cells. The PGBS-COOH presented
large pores, which are also attributed to rearrangement of the polymer chains
during the chemical reactions. The SEM images clearly reveals a plate like
morphology with relatively smooth surface on BS compared to the porous texture
of the PGBS-COOH with rough surfaces. This may account for the higher BET
surface area on PGBS-COOH (110.3 m2/g) than the BS (86.3 m2/g) material. On
examination of the SEM of PGBS-COOH, it is found that polyacrylamide (PAAm)
grafts are deposited on the surface of the unit cells as well as in the intercellular
gaps. The relative smoothness of the BS surfaces are evident, whereas PGBS­
COOH surfaces show more advanced layers of the deposited PAAm grafts on the
surface which reduces the degree of crystallinity. The decrease in crystallinity is
attributed to a reduction in tensile strength of the grafted chain. The PGBS-COOH
presented large pores, which are generated from the rearrangement of the polymer
chain during the chemical reactions.
11{
Chapter 4

Figure 14
Scanning electron micrographs of BS and PGBS-COOH
113
Chapter4

4.4.5 Determination of pHpzc


The surface charge (o-0) of the BS and PGBS-COOH in aqueous phase was
analyzed by the titration method using two different concentrations of electrolyte
(0.01 and 0.001 M NaCl). The pHpzc was determined by plotting cr0 versus system
pH, where cr0 is calculated as per Equation (3 .2). The variation of o-0 with system
pH is shown in Figure 15 and the pHpzc is estimated around 8.0 for BS and 5.5 for
PGBS-COOH. The decrease in pHpzc after surface modification indicates that the
surface becomes more negative after grafting and functionalization. The low pHpzc
of PGBS-COOH indicates that it has more acidic functional groups than the
original BS, or in other words, it is less basic which is advantageous to remove
metal cations through electrostatic interactions. The characteristics of the
PGBS-COOH are: surface area; 110.3 ± 3.8 m2/g; porosity 0.51 ± 0.06 mL/g; pHpzc
5.5 ± 0.3; particle size, 0.096 mm; cation exchange capacity, 2.38 ± 0.12 meq/g;
total acidity, 2.10 ± 0.36 meq/g; carboxylate content, 1.86 ± 0.41 meq/g and
apparent density 0.72 ± 0.08 g/mL.

--- •
25 1.25

...
• --;---+ : O.OIM NaCl
: O.OOIM NaCl
"'a 15 - : PGBS-COOH 0.75 =Q
j
=
CJ

I0 5
:BS
0.25 �
0 i
-5 -0.25 a
tb -15
CJ
fl
-0.75 a...

-25 -1.25
0 2 4 6 8 10 12 14
pH
Figure 15
Effect of pH on the surface charge density of BS and
PGBS-COOH at different ionic strengths
114
Kinetic and Equilibrium Models ...........

4.5 Effect of pH on Metal Adsorption


The pH is the most important factor affecting the adsorption process. To study the
influence of pH on the adsorption capacity of PGBS-COOH, experiments were
performed using various initial pH values ranging from 2.0 to 9.0. The results
obtained are represented in Figures 16-19. As expected, the adsorption of metals
decreases with decreasing pH because the carboxylate functional groups are more
protonated and hence, they are less available to retain the investigated metals. The
binding of metal ions by surface functional groups was strongly pH dependent.
Adsorption begins to show an increasing trend in the pH range of 3.0-4.0 for
Pb(II), Co(II), Hg(II) and Cd(II) and rises steeply upto a pH range of 6.0-8.0.
A maximum adsorption of 95.8 and 85.4 % for Pb(II), 94.0 and 83. 7 % for Co(II)
and 91.3 and 78.0 % for Hg(II) occurred for an initial concentration of 50 and 100
mg/L respectively at a pH of 6.5. A maximum adsorption of 89.1 and 79.8 % for
Cd(II) was observed at an initial concentration of 25 and 50 mg/L respectively. The
removal percentage due to precipitation rises around pH value of 8.0 and after for
all the metal ions is well supported by the respective speciation diagrams. The
maximum adsorption of Pb(II), Co(II), Hg(II) and Cd(II) on BS was found at the
pH range of 5.0-7.0. Below this pH range, a decreasing trend in removal was
observed. The maximum removal capacities onto BS were found to be 36.5 and
25.8 % for Pb(II), 39.5 and 29.5 % for Co(II), 49.4 and 39.0 % for Hg(II)
respectively at an initial concentration of 50 and 100 mg/L. For Cd(II), the
removal was found to be 36.9 and 32.4 % for an initial concentration of 25 and 50
mg/L respectively. It was also noticed that BS is effective within a narrow pH
range of 5.0-6.0. The figures clearly illustrate that the adsorbent PGBS-COOH is
more than twofold superior to BS in adsorbing the tested metals from aqueous
solutions.
115
Chapter4

The distribution and concentration of species in solution depend on the solution


pH. A knowledge of the relationship between the adsorbent surface chemistry and
the species in solution helps to explain the metal sorption mechanisms. The
different species with respect to the metal ions under various pH values are found
to have a +2 oxidation state. The sharp increase in metal uptake on pH 3.0 to 5.5
can be explained by ion exchange. The dependence of metal uptake on pH suggests
that the weak acidic carboxyl groups, R-Coo- (pKa range between 3.5 and 5.5) of
PGBS-COOH are the probable adsorption sites. A perusal of the speciation
diagrams reveals that at a pH of less than 5.5 the predominant metal species
[M2+ and M(OHt] are positively charged (Figures 16-19) and therefore the uptake
+
of metals in the pH below 5.5 is a H - M2+/M(OHt exchange process. At lower
+
pH values the H ions compete with M2+ ions for the exchange sites in the system.
It was also seen that final pH is always less than the initial pH. This indicates that
+
as the metal ions are bound onto the adsorbent, H ions are released from the
-COOH functional groups into the solution leading to a fall in pH. This leads to the
conclusion that PGBS-COOH probably acts as an acid form ion exchanger. The
perusal of speciation diagrams (Knocke and Hemphill, 1981) clearly indicates that
in the range of the highest removal efficiency, the dominant species were M+ and
M(OHt (Baes and Mesmer, 1986). The slight increase in adsorption at higher
pH 7.0-9.0 may be due to the retention of M(OH)2 species in the micropores of the
adsorbent particles.
The pHpzc of PGBS-COOH was found to be 5.5 and above this pH the surface
charge of the adsorbent is negative. Considering that the sorbent consists of
carboxylate groups, which act as an anionic chelating agent, the sorption process
may be attributed to the electrostatic attraction between positively charged metal
ions and negatively charged sorbent. At low pH, H+ ions compete with metal
cations for the surface of the adsorbent, which would hinder metal ions from
116
Kinetic and Equilibrium Models ...........

reaching the binding sites of the adsorbent, caused by repulsive forces. In the
present study maximum adsorption for all the four metal ions occurred at pH
values higher than that of pHpzc ·
The removal of Pb(II), Co(II), Hg(II) and Cd(II) ions by PGBS-COOH may be
represented as
+ +
2 PGBS-COOH + M2 (PGBS-C00)2 M + 2 H (4.9)

PGBS-COOH + M(OHt ��- PGBS-COO M(OH) + Ir (4.10)

where M represents the metals.


Adsorption of metal ions on original BS can be attributed to coulombic
interaction. Coulombic interaction can be reflected from adsorption of cationic
species versus anionic species on adsorbents. BS is basically a lignocellulosic
material. Lignocellulosic materials contain polar functional groups like carboxylic,
phenolic and hydroxylic groups which exist in the cellulose matrix or in the
materials associated with cellulose, such as hemicellulose and lignin. These groups
relatively have strong coulombic adsorption to cations as well as intrinsic
adsorption to anions (Maranon and Sastre, 1991; Ajmal et al., 1996). It is thus
established that graft copolymerization followed by functionalization is an
effective method to improve the adsorption efficiency of BS. As such subsequent
investigations on metal ion adsorption were made only with PGBS-COOH, and an
optimum solution pH of 6.5.
117
Chapter 4

120
(b)
100
Pbz+ Pb(OH)z
.-. 80
-
~
0

r Il
~
'0
60
8-
iZl 40

20

0
(a)
Sorbent dose: 2 gIL
100
Temperature: 300e
PGB8-eOOH
.- Time: 3h
';!. 80
'-"
'a
~ 60
<>+ 100 mgIL
Q
Sl
1:i.A , ... _ #

~ I ...... ·50 mgIL


CI=: 40 I
I I

20

0
0 2 4 6 8 10 12
pH

Figure 16
Effect of pH (a) on the removal of Pb(II) on BS and
PGBS-COOH. [(--) Adsorption, (- - - -) Precipitation]
and (b) on the distribution of Pb(II) species
118
Kinetic and Equilibrium Models ...........

140
(b)
120
2+
Co
_ 100
80
....

� 60

40

20
0
(a)
Sorbent dose: 2 g/L
120
Temperature: 30 °c
-. 100 Time: 3 h
� GBS-COOH

-;
80 <>• 50 mg/L
� l::,,.Ji,. 100 mg/L
a 60
I
�' 100 mg/L
� I --
40

20
"
I 1 50 mg/L

BS
0
0 2 4 6 8 10 12
pH

Figure 17
Effect of pH (a) on the removal of Co(II) on BS
and PGBS-COOH. [(--) Adsorption, (- - - - ) Precipitation]
and (b) on the distribution of Co(II) species
119
Chapter4

120

-
Hg
2+ (b)
100
H g(OH)2

� 80
._
.... 60

00 40

20
0
Sorbent dose : 2 gL (a)

- too Temperature : 30 °c
Time: 3 h

._ 80

60

40

20 50mgL

0
0 2 4 6 8 10 12
pH

Figure 18
Effect of pH (a) on the removal of Hg(II) on BS and
PGBS-COOH. [(--) Adsorption, (- - - - ) Precipitation] and
(b) on the distribution of Hg(II) species
120
Kinetic and Equilibrium Models ...........

120
(b)

-
100

'I. 80

·-
fl.I

y
60 2
Cd(OH)2

i
Cd
40
t
Cd(OH
20

0
(a)

--
Sorbent dose : 2 g,'L
100 Temperature : 30°C
Time: 3h
'I. 80 D • 25 mg,'L
-a... <>+ 50 mg,'L
60
50 mg,'L
� 40
I I
I I
20 I I
., '., 15 m
0
0 2 4 6 8 10
pH

Figure 19
Effect of pH (a) on the removal of Cd(II) on BS and
PGBS-COOH. [(--) Adsorption, (- - - - ) Precipitation]
and (b) on the distribution of Cd(II) species
121
Chapter4

4.6 Effect of Ionic Strength

The heavy metal sorption from real wastewaters is usually carried out in the
presence of various other ions. Ionic strength is one of the functions for expressing
the total amount of coexistent ions. The acid-base and electron-donor properties of
the adsorbent, the flexibility of the polymer chains and the degree of swelling of
the polymer in water solution depend on the ionic strength of the solution. The
effect of ionic strength on the adsorption Pb(II), Co(II), Hg(II) and Cd(II) on
PGBS-COOH was studied at pH 6.5 with different ionic strengths varying from
0.01 to 1.0 M NaCl and at a fixed initial metal concentration of 50 mg/L
(Figure 20). It was found that adsorption of metal ions from the solution decreased
with increase in the concentration of NaCl. The adsorption of metal ions per unit
weight of adsorbent was found to decrease from 24.0 to 17.2 mg/g for Pb(II), 23.7
to 16.7 mg/g for Co(II), 22.9 to 17.2 mg/g for Hg(II) and 19.9 to 14.3 mg/g for
Cd(II) for a change in ionic strength from 0.001 to 1.0 M (Figure 20). Since
considerable change in adsorption is not observed between the ionic strengths
0.001 and 0.01 M, 0.1 M was taken to be the optimum concentration of NaCl for
further studies.
The effect of change in adsorption with increase in NaCl concentration may be
imparted to changes in the properties of electrical double layer or in the metal ion's
activity. Adsorption is sensitive to the change in ionic strength if electrostatic
attraction is a significant mechanism. Thus, here it seems that electrostatic
attraciion plays an important role in the removal of metal ions by PGBS-COOH.
According to the surface chemistry theory developed by Guoy Chapman (Osipow,
1972), when solid adsorbent is in contact with sorbate species in solution, they are
bound to be surrounded by an electrical diffused double layer, the thickness of
which is significantly expanded by the presence of electrolyte. Such expansion
inhibits the adsorbent particles and metal species from approaching each other
122
Kinetic and Equilibrium Models ...........

more closely and, through the decreased electrostatic attraction, leads to the
decreased uptake of metal ions. Lee and Yang (1997) explain the reduction of
metal adsorption by the presence of competing Na+ ions for metal binding
nevertheless, the reduction can also be explained on the basis of different ionic
species present at different chloride concentrations. Perusal of metal speciation
diagram (Knocke and Hemphill, 1981) showed that an increase in chloride
concentration reduces M2+and M(OHt species due to the formation of
chlorocomplexes. The results obtained are in good agreement with the calculated
values for the Langmuir constant 'b' related to binding energy between the cations
and the functional groups of the adsorbent. The results show that, the higher the
stability of the complex formed between the metal ion and the functional groups of
the adsorbent, the weaker the negative influence of the increased ionic strength.

30
Sorbent dose : 2 gL

cb 25 Temperature: 30°C
e Time: 3h
20 pH:6.5
Q
15

10 ••
I :Pb
:Co
e •
Q
:Hg
5 :Cd

0
0 -2 -4 -6 -8
In [NaCl]

Figure 20
Effect of ionic strength on the adsorption of Pb(II), Co(II), Hg(II)
and Cd(II) onto PGBS-COOH
123
Chapter4

4.7 Effect of Contact Time and Initial Concentration


Two of the most important parameters, which define an adsorbent, and evaluate an
adsorption process, are adsorption time and adsorption capacity. Adsorbent time is
particularly important in defining the efficiency of adsorption and it helps to
determine the effluent flow rate through a column for optimum removal of
particular contaminants. Adsorption capacity determines the maximum amount of
metal ions that can be removed per weight of adsorbent. In order to estimate the
adsorption capacity, it is important to allow sufficient time for the experimental
system to reach equilibrium. To establish an appropriate contact time between the
adsorbent and metallic ions solution, the percentage of adsorption of metal ions
was measured as a function of time. The adsorption data for the uptake of Pb(II),
Co(II), Hg(II) and Cd(II) ions by PGBS-COOH with respect to change in contact
time are presented in Figure 21. For this purpose four different initial
concentrations of each metal ions were taken. Batch adsorption experiments were
carried out by shaking 0.1 g of the adsorbent with 50 mL of the metal ion solution
of each concentration at a pH of 6.5 and at room temperature of 30 °c in several
stoppered bottles for different retention times using a temperature controlled water
bath shaker. In each case the limit to contact time is fixed with a view that
consecutive reading would give the same percentage of adsorption. Aliquots of the
aqueous solution were withdrawn at prefixed time intervals and analysed for the
remaining metal concentrations.
It was seen that the adsorption of the tested metals reached equilibrium in
approximately 3 h. It is also noted that the kinetic profile in all the cases exhibited
a relative rapid initial rate of adsorption followed by a slow approach to
equilibrium. The results show that the adsorption of metal ions increases very
rapidly up to about 30-40 minutes occurred by surface reaction and slowly reaches
a constant value beyond which no more of the metal is removed from solution. The
124
Kinetic and Equilibrium Models ...........

characteristic of rapid adsorption is of significant importance in a continuous-flow


metal treatment system since it enables metal uptake over short contact times in the
process (Zhao et al., 1999). A quasi-stationary state obtained within 3 h of shaking
time for all the metal ions was taken as the optimum equilibration period for all the
other experiments.
With change in the concentration of the solution from 25 to 150 mg/L, the
amount adsorbed increased from 12.43 mg/g (99.4 %) to 59.52 mg/g (79.37 %) for
Pb(Il), from 12.35 mg/g (98.8 %) to 56.47 mg/g (75.3 %) for Co(II), from 12.32
mg/g (98.5 %) to 51.89 mg/g (69.2 %) for Hg(II) and from 11.26 mg/g (90.1 %) to
46.34 mg/g (58.2 %) for Cd(II) respectively. Depending upon the initial
concentration, about 50-75 % removal of metals was achieved during the first 30
min of contact time, while only 15-20 % of additional removal occurred in the
following 3 h of contact time. A sharp increase in removal rate at initial stage may
be due to the availability of sufficient number of exchange sites.
The amount adsorbed on the PGBS-COOH at lower metal concentration is
smaller than the amount adsorbed at higher metal concentrations. This is due to the
decrease in the driving force of the concentration gradient, as on decrease with
initial concentration. In addition at higher initial concentration the available site of
adsorption becomes fewer and hence the percentage of adsorption of metal ion
depends on the initial concentration. For a fixed adsorbent dose, the total available
adsorbent sites are limited, thereby adsorbing almost the same amount of sorbate
resulting in a decrease in percentage adsorption corresponding to an increase in
initial sorbate concentration.
125
Chapter4

120
Pb(Il) Co(Il)
100

80

60
• : 25mg'L Sorbent dose : 2 g'L • : 25mg'L
Sorbent dose : 2 lefL
40 + : 50mg'L
° + : 50mg'L Temperature: 30 C
°
,,.... Temperature: 30 C
A : 100mg'L
'1. A : 100mg'L pH: 6.5
� 20 pH: 6.5
• : 150mg'L • : 150mg'L
Agitation: 200rpm Agitation: 200 rpm
0
'120
0
Cd(Il) Hg(Il)
; 100

80

60
• : 25mg'L • : 25mg'L
40 + : 50mg'L + : 50mg'L
A : 100mg'L A : 100 mg'L
20 • : 150mg'L • : 150 mg'L
Agitation: 200 rpm Agitation: 200 rpm
0
0 50 100 150 200 250 300 50 100 150 200 250 300

Time (min)
Figure 21
Effect of contact time and initial concentration on the
adsorption of Pb(II), Co(II), Hg(II) and Cd(II) onto PGBS-COOH

4.8 Adsorption Kinetics

Batch rate analysis can be used to evaluate kinetic and other parameters for the
design of batch mode treatment systems given the desired treatment goals. The
effect of concentration and temperature on reaction rates is well known and
important in understanding reaction mechanism. Several, kinetic models are
available to examine the controlling mechanism of adsorption process and to test
126
Kinetic and Equilibrium Models ...........

the experimental data. The rate constants of adsorption metals from solution were
determined by making use of the pseudo-first-order equation of Lagergren (1898)
and the pseudo-second-order equation by Ho and McKay (2000).
The Lagergren pseudo-first-order equation is represented as

(4.11)

where k1 is the rate constant of pseudo-first-order adsorption, qe and qt denote the


amount of adsorption at equilibrium and at time 't' respectively. The
pseudo-second-order equation is represented by
t 1 t (4.12)
-=--+-
q, k2q; qe
where k2 is the rate constant of sorption (g/mg min), qe and qt are the amount of
metal ions adsorbed (mg/g) at equilibrium and at time 't' respectively. The product
kiq/ is actually the initial sorption rate represented ash= k2q/. Straight line plots
with low correlation coefficients were observed for the Lagergren plots but
correlation coefficients of more than 0.989 were obtained for the straight lines
corresponding to Equation (4.1 2) for adsorption of all the metal ions (figures not
shown).
A comparison between the calculated and observed values of qt versus t for the
adsorption of Pb(Il), Hg(Il), Co(II) and Cd(II) ions onto PGBS-COOH at different
initial concentrations showed good agreement of the pseudo-second-order kinetic
model (Figure 22). This indicates that the adsorption reaction can be approximated
to the pseudo-second-order model. In order to quantitatively compare the
applicability of each model, a normalized standard deviation �q(o/o) is calculated
using the following equation
L{(q/xp _ q/a/ }/q/xp 2
J
6q(%)=IOO (4.13)
n-I
127
Chapter4

where the superscripts 'exp' and 'cal' are the experimental and calculated values
and 'n' is the number of measurements. Kinetic parameters of these models for
different concentrations are calculated from the slope and the intercept of the linear
plots of log(qe-qJ versus t and t!q, versus t and are given in Table 5. The values of
dq(%) and correlation coefficients (r2) are also listed in Table 5. Based on r2 and
dq values it is established that the adsorption of metal ions onto PGBS-COOH can
be best described by the pseudo-second-order kinetic model. The earlier
application of the pseudo-second-order equation to the kinetic studies of
competitive heavy metal adsorption by sphagnum moss peat was undertaken by Ho
et al. (1996). The pseudo-second-order rate expression has also been applied to the
sorption of dyes and organic substances from aqueous solutions (Ho et al., 2001).
The values of qe were found to increase with increase in the initial metal
concentration from 50 to 150 mg/L. The perusal of the data for an equilibrium time
indicates that the metal ion adsorbed qe is higher for Pb(II) followed by Co(II),
Hg(II), and Cd(II) at all concentrations. This is obvious because more efficient
utilization of the adsorptive capacities of the adsorbent is expected due to a greater
driving force by higher concentration gradient pressure. For a decrease in the initial
concentration of metal ions from 150 to 25 mg/L, the values of second-order rate
constant, k were found to increase from 8.97 x 10-3 to 1.35 x 10-3 g/mg min for
2

Pb(II), and from 2.24 x 10-2 to 1.68 x 10-3 g/mg min for Co(II), from 5.73 x 10-3 to
1.23 X I 0-3 g/mg min for Hg(II) and 1.10 X I 0-2 to 2.48 X I 0-3 g/mg min for Cd(II)
respectively. The product kq/ is actually the initial sorption rate represented as
h = kq/. The data clearly show that the initial sorption rate, h is increased with an
increase in initial metal concentration.
128
Kinetic and Equilibrium Models ...........

80 .--�����������--�����������----.
Pb(U) Co(U)
70
60
50
40 -................ .. .......................----
� 30
S 20
� 10 ...... ...

'i �·.::
.. ... ... .. ....
..
80
e : 25mr/L Cd(U) Hg (U)
Sorbent dose:2 rfL
d= 70
Temperature:30°C • : 50 mrfL
e 60 pH:6.5 .i : IOOmr/L
< I : 150mrfL

Agitation : 200rpm
50
40
30 .. . .. . .. . . . . .. . .. . . ..................
20
. . .. ... .. .......... , ......
......................
10 ----------------·······
....
O II"-�"'-�.......�.......�_...�_._�......,:::: -­
"----'�---'---�"--�-'-����
0 50 100 150 200 250 300 50 100 150 200 250 JOO
Time (min)

Figure 22
Effect of contact time on the adsorption of Pb(II), Co(II), Hg(II)
and Cd(II) onto PGBS-COOH and comparison of two kinetic models
with the observed data- (Legends): experimental;{-):
pseudo-second-order; {- - - -): pseudo-first-order
Table 5
Kinetic parameters for the adsorption of Pb(Il), Co(ll), Hg(Il) and Cd(ll) onto
PGBS-COOH at different initial concentrations
Pseudo-first-order Pseudo-second order

Metal Concn q,,exp k2Xl03


k1x102 q,, cal !:iq q,, cal
(mg/L) (mg/g) 1 r2 (g!mg (mg!g
!:iq r2
(min- ) (mg/g) (%) (mg/g) (%)
min) min)

25 12.43 2.66 7.47 67.3 0.942 8.97 1.48 12.83 8.07 0.999
50 23.96 2.26 15.23 64.8 0.957 4.19 2.60 24.89 7.24 0.998
Pb(II)
100 42.62 2.29 30.29 57.4 0.967 2.09 4.18 44.74 10.18 0.999
150 59.52 2.49 40.74 61.5 0.962 1.35 5.29 62.60 12.88 0.997
25 12.35 1.77 3.54 90.5 0.931 22.39 3.49 12.48 5.87 0.998
50 23.50 2.44 11.73 75.2 0.942 6.93 3.92 23.97 7.37 0.999
Co(II)
100 40.56 1.80 23.54 71.9 0.954 2.38 4.17 41.83 5.38 0.999
150 56.47 2.48 37.69 61.4 0.945 1.68 5.75 58.47 5.75 0.999
25 12.32 1.76 8.02 68.7 0.964 5.73 0.93 12.75 14.24 0.998
50 22.88 1.73 16.16 64.4 0.966 2.42 1.42 24.27 14.14 0.999
Hg(II)
100 39.18 1.99 27.42 62.9 0.975 1.60 2.74 41.41 13.18 0.997
150 51.89 1.88 35.36 65.4 0.954 1.23 3.63 54.34 14.45 0.997
25 11.21 2.14 5.69 76.7 0.975 11.01 1.48 11.53 11.67 0.998
50 20.32 2.59 11.19 71.4 0.972 6.13 2.68 20.89 9.62 0.999
Cd(II) n
100 32.67 2.12 16.91 75.9 0.964 3.45 3.92 33.71 12.21 0.999
150 43.64 2.54 26.43 67.8 0.958 2.48 5.08 45.30 8.95 0.999 .... .....
al� N
CD
130
Kinetic and Equilibrium Models ...........

An identical trend in the h and k2 values with the initial concentration has been
reported also by earlier workers who studied the adsorption characteristics of
Cu(II) on chitosan (Wu et al., 2001) and Hg(II) adsorption on sulphurised activated
carbons (Anoopkrishnan et al., 2003). The decrease in the k2 values with increase
in concentration may be due to the progressive decrease in covalent interactions,
relative to electrostatic interactions of the sites with lower affinity for M(II) with an
initial M(II) concentration.
The linearised k2 and Ca values on log-log plots (Figure 23) show high
correlation coefficient greater than 0.971 and the following relationships

k2 == 0.2613 co<-1.3422) (4.14)

k2 == 2.3812 c0<-1.4n1) (4.15)

k2 == 0.0765 co<-0·8382) (4.16)

k2 ==0.1634 co<-1.sus) (4.17)

were obtained for Pb(II), Co(II), Hg(II) and Cd(II) respectively. The decrease in
rate constant with concentration is sharp in the case of Co(II) as indicated by the
large negative value of the slope. The above relationships can be utilized for
predicting the rate constant of adsorption process at any initial concentrations
under similar experimental conditions.
131
Chapter4

-2
Sorbent dose : 2 g'L
-3 Temperature: 30°C
-4 Time: 3h

=
pH :6.5

••
-5
:Pb

..
-6 :Co
-7 :Hg
:Cd
-8
2 3 4 5 6
In C0
Figure 23
Log-log plots of initial concentration and pseudo-second-order
rate constants of the adsorption of Pb(II), Co(II), Hg(II)
and Cd(II) on PGBS-COOH

4.9 Effect of Adsorbent Dose on Co(II) Adsorption

In order to understand the effect of adsorbent dose on Co(II) removal, the


experiments were also conducted by varying the adsorbent dose between 1 and
!Og/L and initial Co(II) concentration between 150 and 600 mg/L at pH 6.5 and
30°C, and in all cases the shape of the removal versus time curves were similar to
that observed in Figure 21. Figure 24 shows the removal percentage of Co(II) as a
function of initial concentration at different adsorbent doses. The percentage
removal of Co(II) increases from 43.4 % (65.1 mg/g) to 99.4 % (14.9 mg/g) by
increasing the adsorbent dose from 1.0 to 10.0 g/L at an initial concentration of
150 mg/L. Similarly at an initial concentration of 600 mg/L Co(II), the percentage
removal increases from 20.2 % (121.2 mg/g) to 87.3 % (52.4 mg/g) by increasing
the adsorbent dose from 1.0 to 10.0 g/L. It was also noticed that the uptake per unit
mass of the adsorbent was higher when its concentration was lower. The reason for
132
Kinetic and Equilibrium Models ...........

such performance of the sorbent, i.e., decrease in unit adsorption with increase in
the dose of adsorbent, may be due to the fact that some adsorption sites remain
unsaturated during the adsorption process, whereas the number of available
adsorption sites increase by increasing the adsorbent dose and that results on the
increase of removal efficiency. Another reason for decrease in unit adsorption may
be the particle interaction such as aggregation, resulting from high adsorbent dose.
Such aggregates would lead to decrease in total surface area of the adsorbent and
increase in diffusional path length.

160 Temperature : 30 C
140 pH :6.5
Agitation time : 3 h
.- 120
'-' 100
t 80
o 60
� 40
20
0
0 200 400 600 800
Co (mg/L)

Figure 24
Effect of adsorbent dose on the adsorption of Co(II) from
aqueous solutions on PGBS-COOH

4.9.1 Modeling of the adsorption curve

The kinetic data obtained from the above experiments were modeled to determine
the percentage removal at any time 't' for the given value of initial concentration
(Co) of Co(II) and adsorbent dose (W,). The experimental kinetic data were fitted
with the relationship of the type
133
Chapter4

t
= a+ ht (4.18)
R

where R is the percentage removal at time t, and a and b are coefficients. From the
slopes and intercepts of the straight line plots of t/R against t for each C0 and Ws, the
corresponding values of a and b were determined (Figure not shown). For each Co
and Ws by fitting equation (4.18) to the experimental data using regression analysis,
it was found that the values of 'a' decreased with the increase of Ws for all values
of Co (Table 6). The average of 'a' values for different initial concentrations with
respect to each adsorbent dose was calculated. The variation of the average values
of 'a' with Ws could be quantitatively represented as

a=---­
c+ dWs (4.19)

where 'c' and 'd' are coefficients. The values of 'c' and 'd' were found to be-9.93
and 9.14 respectively. From Table 6 it is also observed that the constant b showed
an increasing trend with Co values for all values of Ws. A linear regression based on
the data of this study yields a relationship between band Co

b= Co (4.20)
e+ /Co
where e andfare coefficients. The values of e andfwere determined and are given
in Table 7 along with the coefficient of correlation (r2). The data clearly showed a
decreasing trend for 'e' with the increasing Ws values, whereas an increasing trend
for 'f. This shows that values of e and f correlate well with Ws through the
relationship

(4.21)
134
Kinetic and Equilibrium Models ...........

= Ws
f (4.22)
i+ jWS
where g, h, i and} are coefficients. Their values were computed as -4.09 x I0-4,
3.82 x 10-4, 5.50 x 10-2 and 5.74 x 10-3 respectively. Substitution of Equations
(4.19}-{4.22) in Equation (4.18) yields relationship of the type

t I C t
-= +----c----------0 ---------,--
R 9.14ffs -9.93 (4.23)
_ _ _ _
s _
_ _ s
(
-4 W.
+
-2 W.
co
-4.09x10 +3.82xl0 J¥s 5.50x10 +5.74xl0 J¥s . J
-4 -3

which helps to determine the percentage of Co(II) removal at any time for the
given value of Co and Ws .
The above model is used to simulate several variable conditions of Co(II)
adsorption, namely C0, Ws, and pH. (Figure 25). In each case the model is able to
simulate batch data reasonably well for various operational conditions. Further
study is required to more fully understand the above relationship with respect to
real industrial wastewater and other operational conditions. The model can be used
to predict the adsorbent dose required to get the required Co(II) removal at any
given time and initial solute concentration. The required removal efficiency can
also be achieved by controlling the variable such as contact time . and dose of
adsorbent in batch reactor.
135
Chapter4

Table 6
Values of coefficients 'a' and 'b' with respect to coefficient of
correlations (l) for different values of 'Ws' and 'Co'
Initial Co(II) Adsorbent dose, W, (g/L)
conen.,
C0 (mg/L) 1 2 4 8 10
a 0.0560 0.0256 0.0070 0.0043 0.0011
150 b 0.0227 0.0131 0.0107 0.0105 0.0100
r2 0.9990 0.9980 0.9990 0.9990 0.9960
a 0.0885 0.0881 0.0318 0.0147 0.0032
300 b 0.0351 0.0179 0.0115 0.0108 0.0104
r2 0.9960 0.9970 0.9990 0.9950 0.9960
a 0.2090 0.1656 0.0645 0.0178 0.0084
450 b 0.0419 0.0204 0.0130 0.0117 0.0109
r2 0.9960 0.9950 0.9990 0.9990 0.9990
a 0.3328 0.2008 0.0817 0.0369 0.0216
600 b 0.0515 0.0238 0.0133 0.0120 0.0113
r2 0.9950 0.9980 0.9990 0.9980 0.9970

Table 7
Values of coefficients of 'e' and'/' with respect to
coefficient of correlations for different values of 'Ws'
Adsorbent dose, W, (g/L) e f r2
5003.65 13.34 0.998
2 4617.11 30.96 0.987
4 4424.25 68.23 0.996
8 3203.48 78.64 0.997
10 2775.73 84.59 0.999
136
Kinetic and Equilibrium Models ...........

120
+ : Experimental
•• • • •
100 -- : Calculated
80 + : Experimental
-- : Calculated
60

-
Adsorbent dose : 10 g,'L Adsorbent dose : 8 g,'L
40 pH : 6.5 pH :7.5
= 20
Initial concn : 300 mg,'L Initial concn
Temperature
: 600 mg,'L
: 30°C
Temperature : 30°C
120
Adsorbent dose : 2 g,'L
� • : Experimental
� 100 pH : 6.5
-- : Calculated
Initial concn : 530 mg,'L
80 Temperature : 30°C
60
40 • •
Adsorbent dose : 7.5 g,'L
pH :7.0
: 450 mg,'L
Initial concn
+ : Experimental
20 Temperature : 30°C
-- : Calculated
0
0 50 100 150 200 50 100 150 200
Time (min)

Figure 25
Comparison of the experimental and calculated values of Co(II)
removal by PGBS-COOH at different experimental conditions
137
Chapter4

4.10 Effect of Temperature on Adsorption Dynamics

The effect of temperature on the adsorption of metals on PGBS-COOH was studied


at different temperatures of 30, 40, 50 and 60 °c keeping 100 mg/L as the initial
concentration in all cases. Sorption was found to increase with temperature and the
initial sorption rate appeared to be the higher, but the equilibrium reached within 3
h in all cases. This representation of the kinetic curves was confirmed by the order
of magnitude of the kinetic parameters obtained with respect to the two kinetic
equations discussed earlier (Table 8). The plots of tlq, versus time at various
temperatures were found to be linear with high correlation coefficients (Figure 26).
In order to explore the kinetic processes of adsorption, it is a good method of using
different models to fit the experimental data and comparing the normalized
standard deviation, l:!q, using the Equation (4.13). The l:!q (%) values
corresponding to the first-order equation is higher in all cases. The theoretical
values of q, versus time calculated from the two kinetic equations along with
observed values are shown in Figures 27-30. The calculated qe values obtained
from the pseudo-second-order equation agree well with the experimental values.
Based on l:!q (%) and r2 values, it is thus evident that the kinetics of sorption of
metal ions at different temperatures were best described by pseudo-second-order
equation. The pseudo-second-order rate constant values, k2, at different
temperatures along with other kinetic parameters were found to increase with an
increase of temperature, indicating that the adsorption process was endothermic in
nature. For an increase in temperature from 30 to 60 °c, the values of second-order
rate constant, k2 were found to increase from 2.01 x 10·3 to 9.16 x 10·3 g/mg min for
Pb(II), and from 2.33 x 10·2 to 9.13 x 10·3 g/mg min for Co(II), from 1.81 x 10·3
to 7.48 x I 0-3 g/mg min for Hg(II) and 3.58 x I 0-2 to 4.43 x I 0·3 g/mg min for Cd(II)
respectively. The increase in temperature favours the adsorption process by
138
Kinetic and Equilibrium Models .......... .

activating the adsorption sites. The increase in temperature also increases the
chemical forces responsible for adsorption.

6
Pb(II) Sorbent dose : 2 tef1, Co(II)
5 Concentration: 100 mtefl,
pH: 6.5
Agitation: 200 rpm

3
: 30 °C
2 : 40 °C



.@� 1 I : so 0c

-
:60 °C
• ; ·o Cd(II) Hg (II)

i5
4

2
1

0
0 50 100 150 200 50 100 150 200
Time (min)

Figure 26
Pseudo-second-order kinetic plots for the adsorption of
Pb(II), Co(II), Hg(II) and Cd(II) on PGBS-COOH
at different temperatures
139
Chapter4

60
30 °C 40 C
°

so
40
....................
,-.., 30 ................
..............
._, 20
10

60
so
°
60 C so 0c
40
30
_____ ..... ............. ........
20
. _ ............... ..... ............
. .,
10 :
0
0 50 100 150 200 50 100 150 200
Time (min)

Figure 27
Effect of contact time on the adsorption of Pb(II) onto PGBS-COOH at
various temperatures and comparison of two kinetic models with the
observed
. data, (Legends): experimental;(--): pseudo-second-
order; (- - - -): pseudo-first-order
140
Kinetic and Equilibrium Models ...........

Figure 28
Effect of contact time on the adsorption of Co(II) onto PGBS-COOH at
various temperatures and comparison of two kinetic models with the
observed data, (Legends): experimental; (--): pseudo-second-.
order; (- - - -): pseudo-first-order
141
Chapter4

60
30 °C 40 °C
50
40
30 .. , . .. . . . . . . . . . . .
..............
20
a
"Cl 10
,e= 60
,,

"Cl
....
� 60 °C 50 °c
50
=a 40
<
30 . . . . .. . .. . .. . . . . .. . . . .
20 - - - - - - - - - · ·· · · · · · · · · · · ·
, ..
10 ,.
:
0
0 50 100 150 200 50 100 150 200
Time (min)

Figure 29
Effect of contact time on the adsorption of Hg(II) onto PGBS-COOH at
various temperatures and comparison of two kinetic models with the
observed data, (Legends): experimental;(--): pseudo-second­
order;(- - - -): pseudo-first-order
142
Kinetic and Equilibrium Models ......•....

60 ° °
J0 C 40 C
50
40
30
20
---······· -- • . . . - - - - · · ·
. . . . . .. . .. . - - · "' . .. .. . .. .
..
,,
"Cl 10 ,
a.I.
,

-=
Q
60
so 0c
"Cl °
60 C
50
Q
40
<
30
20 · - ... . . . . .. .
- · · · . .. .. . .. .. -- - - · · · · · · ·
_

10 , ,

0
0 50 100 150 200 50 100 150 200
Time (min)
Figure 30
Effect of contact time on the adsorption of Cd(II) onto PGBS-COOH at
various temperatures and comparison of two kinetic models with the
observed data, (Legends): experimental; (--): pseudo-second­
order; (- - - -): pseudo-first-order
Table8
Kinetic parameters for the adsorption of Pb(ll), Co(II), Hg(ll) and Cd(II) onto
PGBS-COOH at different temperatures
Pseudo-first-order Pseudo-second-order

h
Metal Temp qe,exp k1xl02 q., cal llq k2XI03 qe, cal llq
(mg/g) 1 r2 (mg/g r2
(OC) (m in" ) (mg/g) (%) (g/mg (mg/g) (%)
min)
min)
30 42.62 3.46 33.29 57.4 0.990 2.01 4.08 45.09 10.3 0.997
40 43.97 3.34 28.73 60.3 0.986 2.61 5.48 45.79 9.2 0.998
Pb(II)
50 45.93 3.44 22.37 73.7 0.999 4.28 9.49 47.19 6.6 0.999
60 49.13 5.75 19.96 60.5 0.919 9.16 22.57 49.63 3.7 0.999
30 39.45 2.44 23.96 67.8 0.991 2.33 3.99 41.43 13.2 0.995
40 41.66 2.32 22.79 73.2 0.986 2.68 5.00 43.19 12.5 0.996
Co(II)
50 44.79 3.78 22.06 69.1 0.984 4.90 10.13 45.92 5.2 0.999
60 46.38 5.47 18.86 75.9 0.974 9.13 20.24 47.08 2.7 0.993
30 39.31 2.26 27.41 62.0 0.968 1.81 3.07 41.21 10.2 0.999
40 43.18 3.33 30.20 56.4 0.899 2.43 4.85 44.70 6.3 0.999
Hg(II)
50 44.56 4.65 28.45 58.4 0.979 3.94 8.34 46.03 5.4 0.998
60 48.90 4.18 17.44 81.8 0.872 7.48 18.34 49.54 2.9 0.997
30 32.67 1.99 16.48 77.5 0.995 3.58 4.01 33.50 6.2 0.998
40 34.56 2.05 16.07 79.8 0.986 3.90 4.83 35.35 6.6 0.997 (')
Cd(II)
50 36.87 1.16 25.54 78.9 0.953 4.06 5.81 37.85 9.6 0.998
60 38.23 2.37 16.56 80.7 0.968 4.43 6.76 39.07 9.4 0.999 .... .....
� �
ii:. w
144
Kinetic and Equilibrium Models ...........

The influence of temperature on adsorption rate constant can be explained by using


Arrhenius equation.

lnk2 = lnA0 Ea
-- (4.24)
RT
where Ao is the temperature independent factor called frequency factor and Ea is the
activation energy. There is a linear relationship between In k2 and the reciprocal of
absolute temperature 1/T. The values of En calculated from the slope of the plots
and were found to be 42.09, 39.03, 39.56 and 5.71 kJ/mol for Pb(II), Co(II), Hg(II)
and Cd(II) respectively. Energies of activation below 42 kJ/mol generally indicate
diffusion-controlled processes and higher values represent chemical reaction or
surface controlled reaction (Spark, 1995). The relatively low Ea values also suggest
that the sorption has a low potential energy barrier.

4.11 Adsorption Isotherms

The purpose of the adsorption isotherm is to relate the adsorbate concentration in


the bulk and the adsorbed amount at the interface. It maps the distribution of
adsorbate solute between the liquid and the solid phase at various equilibrium
conditions. An adsorption isotherm, besides providing a panorama of the course
taken by the system under study in a concise form, indicates how efficiently an
adsorbent will take up and allows an estimate of the economic feasibility of the
adsorbent's commercial application for specified solutes. The analysis of the
isotherm data is important to develop an equation which accurately represents the
results and which could be used for design purpose. Figure 31 shows the
experimental sorption isotherm for four metal ions on PGBS-COOH at different
temperatures from 30 to 60 °c. Initially the adsorption is quite rapid, which is
followed by a slow approach to equilibrium at high metal concentration. It is also
145
Chapter4

observed that for the same equilibrium time, the adsorption is higher for greater
values of initial metal ion concentration. The curvature of the isothe1m decreased
as the Ce values increased considerably for a small increase in qe , The L-type
nature of the curve obtained in the present investigation is favourable for
adsorption and indicates a strong tendency in the process of monolayer formation
(Giles et al., 1960). The normal shape of the isotherm for all metals represent the
formation of a monolayer on the PGBS-COOH surface, meaning that the adsorbent
is favourable at low concentration range as indicated by the sharp line (Faust and
Aly, 1987). To determine the adsorption capacity of the PGBS-COOH, a study of
adsorption isotherm was attempted by analyzing the equilibrium data by a
computer simulation technique to fit the different isotherm models.
The Langmuir, Freundlich and Scatchard isotherm models were used to
interpret equilibrium isotherm data. The Langmuir isotherm is based upon an
assumption of monolayer adsorption onto a surface containing a finite number of
adsorption sites of uniform energies of adsorption with no transmigration of
adsorbate in the plane of the surface. The adsorbed layer is thus one molecule in
thickness and all sites are equal, resulting in equal energies and enthalpies of
adsorption. The strength of the intermolecular attractive forces is believed to fall
off rapidly with distance. The Langmuir isotherm is given by

(4.25)

The sorption data were analyzed according to the linear form of the Langmuir
isotherm

(4.26)
146
Kinetic and Equilibrium Models ...........

where Q0 is the amount of adsorbate required to form a monolayer coverage of


available adsorption sites, where qe is the absorbed equilibrium metal concentration
on PGBS-COOH (mg/g) and Ce is the equilibrium solution phase metal
concentration in mg/L and b is the equilibrium constant (L/mg).
The Freundlich model is one of the most widely used isotherms for the
description of adsorption equilibrium. Freundlich isotherm can describe
satisfactorily the adsorption of organic and inorganic compounds on a wide variety
of adsorbents including activated carbons and synthetic resins. Furthermore, the
isotherms can be incorporated readily into the adsorption modeling for process
design. The Freundlich isotherm model assumes heterogeneous surface energies in
which energy term in Langmuir equation varies as a function of surface coverage.
The Freundlich isotherm equation is given as
qe = Kf. + cefn (4.27)

A linear form of this equation is commonly used by taking logarithms on both sides
and is represented as
1
log qe = log KF +-log c. (4.28)

where KF and 1/n are Freundlich isotherm constants related to adsorption capacity
and intensity of adsorption respectively. This isotherm is another form of the
Langmuir approach for adsorption on an "amorphous" surface. The amount
adsorbed is the summation of adsorption on all sites, each having a bond energy.
The · Freundlich isotherm is derived by assuming an exponential decay energy
distribution function . inserted in the Langmuir equation. It describes reversible
adsorption and is not restricted to the formation of the mono layer. The Scatchard
isotherm represents intermediate situations, close to the Langmuir model. The
Scatchard isotherm equation for a linear plot is represented by the following
equation
147
Chapter4

(4.29)
�e = Ks(Q/ -qe)
e
where Qs° and Ks are Scatchard constants related to the maximum monolayer
adsorption capacity and equilibrium adsorption constant respectively.
Linear regression was frequently used to determine the most fitted model
throughout the years and the method of least squares has been frequently used for
finding the parameters of the models. However Ho (2004c) indicated that
transformations of non-linear isotherm equations to linear forms implicitly alter
their error structure and violate error variance and normality assumptions of least
squares. The isotherm constants of all the previous equations are very useful
parameters for predicting adsorption capacities and also for incorporating into mass
transfer relationships to foresee the design of batch reactors. The values of all the
isotherm constants were calculated from the slope and intercept of the plots using
regression analysis and are shown in Table 9. In this manner, the experimental and
theoretical values of q are shown in Figure 31. It can be seen that, the experimental
data presented a good fitting in relation to the Langmuir adsorption model for all
the metals. The validity of the isotherm models is tested by comparing the
experimental and calculated data using the normalized standard deviation liq (%)
represented by the Equation (4.13). The values of liq(%) calculated are included in
Table 9.
The plots of the experimental qe and Ce values as specific sorption (C/qe)
against the equilibrium concentration (Ce) for the metal-PGBS-COOH systems are
shown in Figure 32. The plots of C/qe versus Ce are linear over the whole
concentration range studied showing the applicability of Langmuir isotherm. This
is further confirmed by the fact that the correlation coefficients (r2) for almost all
the experimental systems are all greater than 0.98 as shown in Table 9. The values
of(! and b for the removal of metals by PGBS-COOH were determined from the
148
Kinetic and Equilibrium Models ...........

slopes and intercepts of the plots using regression analysis. The values of Q0 and b
were found to be 185.34 mg/g and 0.0167 L/mg for Pb(II), 166.93 mg/g and 0.0093
L/mg for Co(II); 137.89 mg/g and 0.0092 L/mg for Hg(II), 65.89 mg/g
0.0341 L/mg for Cd(II) respectively at 30 °c. The adsorption capacity with respect
to all the metal ions increased with increase in temperature. On the other hand, the
equilibrium adsorption, qeq was smaller than Q0• This showed that the adsorption of
metal ions on PGBS-COOH is a monolayer type in which the surface of the
adsorbent is not fully covered (Ozer et al., 1999).
The experimental results showed that the adsorption capacity of the sorbent for
the various metal ions decreased in the order Pb> Hg> Cd at 30°C. The observed
order of uptake for metal ions was the same as that of their increasing ionic radii,
i.e., their decreasing hydrated ionic radii. The hydrated radii of the metal ions
+
studied are: 0.401 for Pb2\ 0.413 for Hg2+ and 0.426 for Cd2 . The smaller the
hydrated ionic radius, the greater its affinity to penetrate into smaller pores and,
thus, have greater access to active groups of the adsorbent. This suggests that the
energy required for the dehydration of the metal ion, so that they could occupy a
site in the adsorbent, plays an important role in determining the selectivity for the
metal ion. But Co(II) with a larger hydrated ionic radii was found to have been
removed more efficiently than either Hg(II) or Cd(II) indicating the fact that Co(II)
might have been coordinated to the various coordinating sites other than the
-COOH groups of the adsorbent. The presence of lateral interactions in the case of
Co(II) adsorption is evidenced by the low l}.q % values with respect to the
Freundlich isotherm. Since the results show that adsorption capacity is not equal
for all studied cations, it seems that PGBS-COOH does not behave unique
alongside of various cations. This proposes that all active groups in PGBS-COOH
are not available to all cations or specific interactions other than electrostatic
interactions occur in adsorption process between the PGBS-COOH and cations in
149
Chapter4

solution. These interactions may be complexation, adsorption, ion exchange, size


exclusion etc. The relative preference for these cations may also be explained by
the higher stability of the complex formed by Pb(II) and Co(II) with carboxylate
groups in PGBS-COOH compared with those formed by Hg(II) and Cd(II) (Manju
et al., 2002).
The values of KF and 1/n were calculated from the intercept and the slope of
the Freundlich plots and are included in Table 9. Examination of the data shows
that the Freundlich isotherm also provides a good description of data over the
concentration range studied for the metal ions even though the correlation
coefficients are comparatively less except for Co(II). Adsorption capacity of the
metal onto the adsorbent represented by KF, is found to increase with temperature
in all cases. The ultimate adsorption capacity of the adsorbent can be calculated by
substituting the required equilibrium concentration in the Freundlich equation.
Thus, for an equilibrium concentration of 1.0 mg/L of metal ions, 1.0 g of
PGBS-COOH can remove 14.33, 10.49, 5.28 and 6.64 mg of Pb(II), Co(II), Hg(II)
and Cd(II) at 30 °c. The values of 1/n lie between O and I and thus indicate good
adsorption of metal ions onto PGBS-COOH. The forces between the surface layer
are attractive if n is less than unity and repulsive if n is greater than unity.
150
Kinetic and Equilibrium Models ...........

250 200
Pb(II) , Co(II)
,,. .... ,-
200 , .... .... --
-:-.;� 150
150
100
100

-
: Langmuir Sorbent dose : 2 giL
50 Time : 3h
50 : Scatchard
pH: 6.5
: Freundlich
0 0
� 0 100 200 300 400 500 0 100 200 300 400 500 600
100 250
Cd(II) Hg(ll)
200
75
150
50
100
:40 °C
25
• : 50 °C 50
• : 60 °C
0 0
0 100 200 300 400 0 200 400 600.

Ce (mg/L)

Figure 31
Comparison of the model fit of various isotherms to the observed
isotherm data for the adsorption of Pb(II}, Co(II}, Hg(II) and Cd(II) onto
PGBS-COOH at different temperatures. (Legends): experimental; (-):
Langmuir; (- - - -): Scatchard; (- - -): Freundlich
151
Chapter4

Figure 32
Langmuir isotherm plots for the adsorption of Pb(II), Co(II), Hg(II) and
Cd(II) onto PGBS-COOH
Table 9 ....~ ~

C1I

Langmuir, Freundlich and Scatchard isotherm constantsfor the adsorption of Ph(/I), Cor/I), =
tD
::r.
N

n
Hg(!/) and Cd(/I) onto PGBS-COOH ~

Langmuir constants Freundlich constants Scatchard constants =


~
Temperature rrJ
.c
(f Q/
-............
b /),.q /),.q ,; /),.q
(mg/g) (Llmg)
? KF lIn ? (mg/g)
ksx 103 (%)
c::
(%) (%)
3D 185.34 0.985 0-
0.0167 0.999 2.7 14.33 0.426 0.944 11.5 187.02 16.40 3.3 '"t
c::
40 191.52 0.0226 0.999 4.9 19.79 0.384 0.958 9.9 187.53 24.13 0.976 6.3 !3
Pb(II)
50 194.34 0.0342 0.999 8.2 26.69 0.347 0.956 10.4 191.38 35.30 0.968 8.4 ~
0
60
30
204.97
166.93
0.0465
0.0093
0.999
0.970
4.1
7.7
33.28
10.49
0.326
0.424
0.966
0.989
9.6
5.3
204.09
175.77
43.86
7.50
0.981
0.908
21.6
47.4
-
~
tD
{ /l

40 169.77 0.0129 0.981 7.3 13.54 0.400 0.996 4.0 170.46 11.88 0.895 36.9
Co(II)
50 172.66 0.0184 0.990 7.3 17.92 0.368 0.995 3.5 169.75 18.03 0.901 36.8
60 181.46 0.0250 0.993 7.8 22.10 0.353 0.988 5.2 175.28 26.85 0.905 30.4
30 137.89 0.0092 0.998 4.9 5.28 0.520 0.924 14.7 150.49 7.53 0.909 5.6

Hg(II) 40 161.74 0.0120 0.991 4.4 7.60 0.501 0.918 15.5 171.53 10.54 0.955 5.2
50 191.03 0.0170 0.992 6.2 11.92 0.471 0.947 13.2 210.74 12.63 0.932 10.4
60 210.50 0.0294 0.997 6.4 20.72 0.412 0.957 11.2 213.56 26.80 0.917 13.3
30 65.89 0.0341 0.998 4.2 6.64 0.424 0.941 14.3 64.45 35.97 0.975 5.2
40 68.77 0.0382 0.999 3.1 7.42 0.416 0.945 14.1 67.70 38.83 0.985 6.6
ed(II)
50 76.65 0.0401 0.998 5.7 8.52 0.413 0.959 12.5 74.14 43.13 0.961 11.8
60 84.22 0.0440 0.997 6.2 9.49 0.417 0.968 12.6 83.24 42.94 0.952 22.9
153
Chapter4

The Scatchard isotherm represents intermediate situations, close to the Langmuir


model. The Scatchard isotherm equation for a linear plot is represented by the
following equation

�e =Ks(Q/-qeJ (4.30)
e

where Qs° and Ks are Scatchard constants related to the maximum monolayer
adsorption capacity and equilibrium adsorption constant respectively. The Qs°
values obtained in all cases are very much close to the corresponding Langmuir
monolayer capacities.
Temperature has a direct influence on the amount of the sorbed substance. In
the present investigation, the adsorption experiments were performed in the
temperaturt;; range of 30-60 °c. It was found that, according to the adsorption
isotherm, the amount of metal adsorbed on PGBS-COOH increase with the
solution temperature. It is observed that the values of Q0 increase with increase of
temperature. The Q0 values increased from 185.34 to 204.97 mg/g for Pb(II) and
from 166.93 to 181.46 mg/g for Co(II) and from 17.89 to 210.50 mg/g for Hg(II)
and from 65.89 to 84.22 mg/g for Cd(II) with a rise in temperature from 30 to
60 °c indicative of the endothermic nature of adsorption. In general, the formation
of adsorption complexes between transition metal cations and carboxylated ligands
is endothermic, whereas amine ligands complex formation is exothem1ic (Volesky,
1990). Since the temperature of most real wastewater was higher rhan room
temperature, PGBS-COOH had an advantage over those adsorbents whose uptake
capacity decreases with increase in temperature.
The results also show that b is an increasing function of the solution
temperature in all cases. The increase of b with temperature indicates that the
affinity for metal ions is favoured by high temperature, and therefore, this
154
Kinetic and Equilibrium Models ...........

adsorption process is endothermic in nature. The increase in capacity with respect


to temperature indicates that some kind of chemical interaction may be taking place
during adsorption process. In order to assess the different isotherms and their
validity to correlate experimental results, the theoretical plots for each isotherm
have also been shown with the experimental data for the adsorption of the M(II)
ions in Figure 31. Langmuir equation gives best fit over the entire range of
concentrations for all the ions.
The experimental results showed that the adsorption capacity of the adsorbent
for the various metal ions expressed on a weight basis, decrease in the order

Pb(II) > Co(II) > Hg(II) > Cd(II) at 30 °c.

However, when the uptake was expressed on a molar basis the following order
was obtained.

Co(II) > Pb(II) > Hg(II) > Cd(IJ).

Certain metal ions bind with a small proportion of functional groups on


PGBS-COOH more strongly than others. This leads to competitive effects which
may not be explained in terms of free energies of adsorption but possibly by
preferential adsorption of different metals at different high energy sites that
represents a small fraction of the total site density. To explain the preferential
sorption of the metals onto the adsorbent the predicted order of bonding preference
based on covalent index should be Hg(II) > Pb(II) > Cd(II) > Co(II), while on the
basis of ionic index, the strongest bond should be formed by the metal with the
greatest charge-to-radius ratio requiring the order Co(II) > Pb(II) > Cd(II) > Hg(II)
(Stumm and Morgan, 1996). The PGBS-COOH adsorbed Pb(II) and Co(II)
strongly suggesting that neither a purely electrostatic nor a covalent picture of
surface bonding is adequate. As Pb(II), Co(II) and Hg(II) happen to be the most
155
Chapter4

easily hydrolysed metals, it might be suggested that adsorption and hydrolysis are
correlated in same way. No other specific reason can at present be attributed to the
observed fact regarding the higher adsorption capacity of Co(II).

4.12 Thermodynamic Parameters

Thermodynamic parameters such as standard Gibbs free energy change (!lG°),


enthalpy change (llI-f\ and entropy change (M1) of the adsorption of M(II) with
PGBS-COOH were calculated using the following equations and are presented in
Table 12.
!lG0 =-RT In b 1 (4.31)

l'lS o - l'lH o
In b, = (4.32)
R RT

where b 1 is the Langmuir constant when concentration terms are expressed in


mol/L. As a result, it can be said that the M(II) adsorption shows clearly that the
adsorption process is composed of two contributions, enthalpic and entropic, which
characterize whether the reaction is spontaneous. The van 't Hoff plots In b 1 versus
I/Twas found to be linear (Figure 33). The values of !!JI, as calculated from the
slope of the plot by regression method were found to be 29.17, 27.81, 32.13 and
7.15 kJ/mol, for Pb(II), Co(II), Hg(II) and Cd(II) respectively. This suggests the
endothermic nature of adsorption of metal ions onto PGBS-COOH. This is in
accordance with the finding that when a divalent ion binds on the adsorbent, !!JI
will be positive (Biskup and Subotic, 1998) when 2 rH > rA, i.e., 2 rtt > rM where rH
= 0.154 nm is the ionic radius of the Ir and rM is the ionic radius of the M2+. The
values of rM are 0.119, 0.078, 0.112 and 0.095 nm for Pb2\ Co2+, Hg2+ and Cd2+
(Greenwood and Earnshaw, 1998). The values of !lG0 (Table 10) are negative and
small which is an indication of the feasibility of the process and the spontaneous
156
Kinetic and Equilibrium Models ...........

nature of sorption. Further, the value of �G0 becomes more negative with rise in
temperature showing increase in feasibility of adsorption at higher temperatures.
The values of �s1 were found to be 163.84, 144.16, 167.91 and 92.20 J/mol/K
respectively for Pb(II), Co(II), Hg(II) and Cd(II). The positive values of� show
the increase in randomness at the solid-solution interface during adsorption. The
relatively small positive value in the system under investigation indicates that as a
result of adsorption no significant structural change occurs in the adsorbent
material.

10.0

9.0

-:c 8.0

7.0 �
•:P
+ :Co
6.0 • :Hg
e :Cd
5.0
0.0029 0.003 0.0031 0.0032 0.0033 0.0034

1/f (K-1)
Figure 33
Plots of In b 1 for the adsorption of Pb(II), Co(II}, Hg(II) and Cd(II)
as a function of 1/T

The effect of isotherm shape can be used to predict whether a sorption system
is favourable or unfavourable in batch process. According to Hall et al. (1966) the
essential features of the Langmuir isotherm can be expressed in terms of separation
factor or equilibrium parameter RL which can be calculated from the relationship.

R - I (4.33)
L - 1 +bC0
157
Chapter4

where Co is the initial concentration (mg/L) and b is the Langmuir constant. RL is a


dimensionless separation factor, indicating the shape of the isotherm. The isotherm
is unfavourable when RL > I, the isotherm is linear when RL = I, the isotherm is
favourable when O < R1, < l and the isotherm is irreversible when R1, = 0.
For all the four metal ions, the RL values corresponding to four different
concentrations viz., 150, 200, 300 and 400 mg/L at four different temperatures
were computed. Increase in temperature decreased the RL values in all cases. It can
be said that adsorption of M(II) onto PGBS-COOH is most favourable at 60 °c. It
is also observed that the RL values decreased with increase in initial concentration
of metal ions. This indicates that sorption is more favourable for higher initial
concentrations than for the lower ones. A maximum value of 0.4212 is computed
for Hg(Il) with an initial concentration of 150 mg/L at 30 °c and a minimum value
of 0.0542 is computed for Cd(II) at 60 °c. The values of RL shown in Table 10
indicate that the adsorption of metals on PGBS-COOH is a favourable process
because RL values lie between O and I.

4.13 lsosteric Heat of Adsorption

The binding sites on an adsorbent surface are either homogeneous or


heterogeneous, based on the binding energy level of the sites. For the homogeneous
sites binding model, the surface sites possess equal intrinsic binding energy for
which the isosteric heat of adsorption would be independent of adsorption densities
(Mohapatra et al., 2004). Information with reference to the magnitude of the heat
of adsorption and its variation with surface coverage can provide useful indication
regarding the nature of the surface and the adsorbed phase. The heat of adsorption
determined at a constant amount of sorbate adsorbed is known as the isosteric heat
of adsorption (11llx), The magnitude of 11llx is calculated by ineans of the Clausius­
Clapeyron equation.
158
Kinetic and Equilibrium Models ...........

Table 10
Equilibrium and thermodynamic parameters/or the adsorption of
Pb(Il), Co(Il), Hg(Il) and Cd(ll) onto PGBS-COOH
Cl) RL
Ele,,...., values
t,.G
O
t,,Jf l:!.S1
Cl) Cl)u
c..o
� sCl) '-' 150
mg/L
200
mg/L
300
mg/L
400
mg/L
(kJ/mol) (kJ/mol) (J/mol/K)
E-<

30 0.2851 0.2303 0.1664 0.1307 -20.53


,,...., 40 0.2286 0.1812 0.1293 0.1001 -21.99
I::::
29.17 163.84
50 0.1631 0.1281 0.0893 0.0685 -23.80
60 0.1254 0.0973 0.0671 0.0515 -25.39
30 0.4173 0.3494 0.2637 0.2117 -15.89

Q 40 0.3395 0.2783 0.2045 0.1616 -17.28


27.81 144.16
50 0.2665 0.2141 0.1537 0.1199 -18.76

60 0.2103 0.1664 0.1175 0.0908 -20.20

-
30 0.4212 0.3531 0.2668 0.2144 -18.93
,,...., 40 0.3577 0.2946 0.2178 0.1728 -20.26
'-'
32.13 167.91
50 0.2816 0.2272 0.1639 0.1282 -21.85
60 0.1848 0.1453 0.1018 0.0784 -24.04
30 0.1641 0.1291 0.0931 0.0691 -20.77
40 0.1503 0.1172 0.0812 0.0621 -21.74
7.15 92.20
,,...., 50 O.l431 0.1114 0.0771 0.0592 -22.58
'-'

u 60 0.1312 0.1022 0.0708 0.0542 -23.56


159
Chapter4

L1Hx (4.34)
= --- 2
RT
The equilibrium concentration Ce, at constant amount of metal adsorbed, obtained
from the adsorption isotherm data at different temperatures are used for the plots of
In Ce versus 1/Tfor different amounts of metal adsorption (Figure 34). The plots of
In Ce versus 1/T were found to be linear with good correlation and the values of
t:Jlx,

•••
7.0 Surface loading (mg/g) Pb(Il) Surface loading (mg/g) Co(II)

6.0
155
150 • 140
5.0 • 145
140
0 135

A 130

• 115
120

4.0
tJ. 130 0 110
D 125 /J. 105
0 115 D 100
3.0

-= 7.0

•••
Cd(II) Surface loading (mg/g) Hg(II)

•••
Surface loading (mg/g)
6.0 58 108.0
56 101.5
5.0
0
• 54
52
50
• 97.5
90.0
0 85.0
4.0
tJ. 48 /J. 80.5
D 46 D 63.5
3.0 0 44

2.8 3.0 3.2 3.4 2.8 3.0 3.2 3.4

1/f (K-1) X 103


Figure 34
Plots of In Ce for Pb(II), Co(II), Hg(II) and Cd(II)
at constant amount adsorbed as a function of 1/T
160
Kinetic and Equilibrium Models ...........

100
Pb(II) Co(II)
80

- .......
60
40
20
...... ... -+

=..
100 120 140 160 18075 100 125 150
100
Cd(II) Hg(II)
80
60
��
40
20
0
40 50 60 70 50 75 100 125 150
Surface loading (mg/g)

Figure 35
Variation of AHx with respect to surface loading

were calculated from the slopes of plots. The average values of !)J!x were found to
be 49.89, 32.39, 28.40 and 27.98 kJ/mol respectively for Hg(II), Pb(II), Co(II) and
Cd(II) and are independent of surface loading (Figure 35). This points out that the
surface of the adsorbent is energetically homogeneous and lateral interaction
between adsorbed metal ions is nearly absent (Wu et al., 200 l ).
161
Chapter4

4.14 Comparison with Other Adsorbents

The removal of metals by weak carboxylic cation exchanger such as Amberlite


IRC-64, Amberlite IRC-88, Amberlite CG-50 and Duolite ES-468 has been
reported. All these adsorbent materials are very expensive and are available for
$ 190-390 per kilogram g of resin. Ceralite IRC-50 (similar to Amberlite IRC-50)
produced by Central Drug House, Bombay, India is the cheapest variety of
commercial weak cation exchanger available in India which is based on
polystyrene-divinylbenzene copolymer having carboxylate functionality. The cost
of this material is$ 70 per kilogram. Adsorption isotherm study was conducted to
determine the adsorption capacity of Ceralite IRC-50. The experimental isotherm
data are shown in Figure 36. Equilibrium isotherm data at pH 6.0 and temperature
30 °c were also correlated using Langmuir, Freundlich and Scatchard isotherm
equations. The Langmuir, Freundlich and Scatchard isotherm constants for the
adsorption of M(II) ions on Ceralite IRC-50 were calculated using regression
analysis and are summarized in Table 11. In order to compare the different
isotherms and their ability to fit in the experimental data, the computed data for
these isotherms have been shown with experimental data for the adsorption of
M{II) (Figure 36). The results from Table 11 show that �q % obtained for
Freundlich and Scatchard isotherms are larger. Hence for Pb(II), Co(II), Hg(II) and
Cd(II) adsorption on Ceralite IRC-50, the Langmuir isotherm best fits the
experimental data in the entire range of concentration. In addition the correlation
coefficient (r2) for the Langmuir isotherm is high. The maximum adsorption
capacity Q' for Ceralite IRC-50 calculated from the Langmuir plot were found to
be 122.69, 58.81, 102.59 and 46.71 mg/g for Pb(II), Co(II), Hg(U) and Cd(II)
respectively which are considerably lower than those calculated for PGBS-COOH
(Table 11).
��
::S N

::r.
tD
n

&
-
t':r1
..c
e.....
a"
::l.
§
Table 11

Langmuir, Freundlich and Scatchard isotherm constants for the adsorption of Pb(II), Co(Il), ,:i..
I!..
Hg(ll) and Cd(ll) onto Ceralite IRC-50 at 3ffC a,

Langmuir constants Freundlich constants Scatchard constants


Metal b tiq
(! r2 &/
KF r2 &/ Qs° 3
r2
! %l !%l !mlifll !%l
1/n ksx 10
!m� !Llmfll
Pb(II) 122.69 0.0149 0.999 7.7 10.11 0.404 0.987 18.3 130.89 1.34 0.987 12.5
Co(II) 58.81 0.0174 0.999 3.2 3.54 0.343 0.954 45.9 64.87 1.18 0.991 5.33
Hg(II) 102.59 0.0207 0.999 5.3 11.55 0.357 0.978 14.5 104.36 2.05 0.983 9.7
Cd(II) 46.71 0.0305 0.999 8.9 5.08 0.398 0.984 11.2 47.76 2.73 0.979 12.4
163
Chapter4

200 00
Pb(II) Hg(II)

150
_ .... _ ....
150

100
-�.-:;,:- ....... 100 ...-:'-; - --- -.-
so so

"""r
,-..,
� 0 0
0 200 400 600 0 200 400 600
� 70 70
Cd(II) Sorbent dose : 2 glL Co(D)
60 60
.. .. . . . .. .. - ·
-- _
Time : 3h

-- --
so pH :6.5 50

--- -- .... --
....
. .. . .. . -
40 40
30 + : Experimental 30
20 : Langmuir 20
: Scatchard
10 : Freundlich 10
0 0
0 100 200 300 400 0 200 400 600 800
Ce (mg/L)

Figure 36
Comparison of the model fit of various isotherms to the observed
isotherm data for the adsorption of Pb(II), Co(II), Hg(II) and Cd(II) onto
Ceralite IRC-50 at 30°C
164
Kinetic and Equilibrium Models ...........

Table 12
Come_arison ol adsore_tion cae_acities (fl) with other adsorbents
Metal Adsorbent (! References
(m��)
Pb(II) Tea leaves 150.3 Singh et al, (1993)
Flax shive carbon 147.1 El-Shafey et al. (2002)
Condensed tannin 114.9 Zhan and Zhao (2003)
Marine algae 85.7 Jalali et al. (2002)
Sawdust 22.2 Taty-Costodes et al, (2003)
PGBS-COOH 185.3 Present work
Co(II) Soil 1.5 Rawat et al. (1996)
Styrene-g-polyethylene
25.0 Choi and Nho (1999)
membrane
Zeolite 85.0 Ebner et (ll. (2001)
Magnetite-silicate
9.4 Ebner et al. (2001)
composite
Cation exchange resin
86.2 Rengaraj and Moon (2002)
IRN 77
PGBS-COOH 166.7 Present work
Namasivayam and Periasamy
Hg(II) Pea nut hull carbon 109.9
(1993)
Sago waste activated
55.6 Kadirvelu et al. (2004)
carbon
Gebremedhin-Haile et al.
Modified zeolite 10.21
(2003)
PGBS-COOH 137.9 Present work
Cd(II) Flax shive carbon 39.1 El-Shafey et al. (2002)
Sawdust 19.1 Taty-Costodes et al. (2003)
Straw activated carbon 11.1 Larsen and Schierup (1981)
Bagasse fly ash 1.2 Gupta et al. (2003)
PGBS-COOH 65.9 Present work
165
Chapter4

In order to justify the validity of a treatment process, the adsorptive capacity of


the PGBS-COOH must be compared with other sorbents examined for the
treatment of M(II) under similar conditions. The maximum adsorption capacity, Q0
of the PGBS-COOH used in the present study at 30 °c was found to be very much
higher than the values reported in the literature (Table 12). The high values of Q0
for PGBS-COOH are diagnostic of the higher capacity of the adsorbent.

4.15 Cost Estimation

Cost is an important parameter for comparing the adsorbent materials. In general,


an adsorbent can be assumed as "low cost" if it requires very little processing, is
abundant in nature, or is a byproduct or waste material from another industry. The
cost-effective and economic removal of toxic heavy metals from sewage, industrial
and mining wastewaters can be done only with low-cost and easily available
adsorbents. The raw material used in the present study, banana stalk, is widely
available free of cost, as it is an agricultural waste. After considering the expenses
for transport, chemicals, electrical energy etc., the final product would cost
approximately $ 105 per ton. The relative cost of the final product (PGBS-COOH)
will be lower than that of commercially available polymeric cation exchangers,
available in India, such as Ceralite, Duolite and Amberlite. The comparatively
lower cost and good adsorption capacity of PGBS-COOH indicate its
complimentary use as low-cost adsorbent for wastewater treatment.

4.16 Wastewater Treatment

4.16.1 Testing with industrial wastewater containing Hg(II)

The utility of the adsorbent material has been demonstrated by treating


with wastewater from Chlor-alkali industry situated in Cochin (India). Industrial
166
Kinetic and Equilibrium Models ...........

wastewater samples collected industry were characterized by standard methods


(APHA, 1992). Composition of wastewater samples is given in Table 13. Since
the amount of Hg(II) in wastewater sample 1 was found to be very low (2.86
mg/L), it was spiked with a Hg(II) solution so that the final concentration of
Hg(II) was 50 mg/L. Sample 2 was used as such without altering the
composition. The typical concentration of Hg(II) in Chlor-alkali industry
wastewater is in the range of 7.0-35.0 mg/L (Data obtained through personal
communication from a chief chemist of the industry). The effect of adsorbent
dose on Hg(II) removal by PGBS-COOH is given in Figure 37. Almost
complete (100%) removal of Hg(II) from 50 mL wastewater was possible with
125 mg (2.5 g/L) and 80 mg (1.6 g/L) of the adsorbent dose for Samples l and
2 respectively which is in good agreement with that obtained from the batch
experiments mentioned above. Examination of the plotted curves in Figure 37
also reveals that at a fixed Hg(II) concentration, the percentage adsorption
increases with increasing adsorbent dose however, unit adsorption or adsorption
per unit mass of adsorbent decreases, the effect being more pronounced the
higher the adsorbent dose. Adsorption of Hg(II) per unit weight of adsorbent
decreases from 57.50 mg/g (46.0 %) at 0.4 g/L of adsorbent to 20.0 mg/g (100
%) at 2.5 g/L for Sample 1 and from 46.41 mg/g (78.0 %) at 0.4 g/L adsorbent
to 14.88 mg/g (100 %) at 1.6 g/L for Sample 2. This is due to the fact that as
the dose of adsorbent (g/L) is increased, there is less commensurate increase in
adsorption resulting from the lower adsorptive capacity utilization of adsorbent.
The results of the experiment can be used to develop a mathematical
relationship between percentage removal (R) and adsorbent dose (ms) [using
boundary conditions]. This relationship for which the correlation coefficient
(r2) > 0.99 is
167
Chapter4

R = ms for Sample 1 (4.35)


(5.80 X 10- ) + (7.56 X 10-3) ms
3

R = ms for Sample 2 (4.36)


(1.70 X 10- )+ (9.04 X 10-3 ) ms
3

Table 13
Composition of ch/or-alkali industry wastewater
Composition (mg/L)
Hg: 50; Pb: 2.7; Cd: 0.5; Mg: 25.6; Ca: 41.2; Na:
Sample I 280.8; P04 : 10.9; N03 : 16.5; NH4 : 20.7; Cl: 398.39;
BOD: 58.4; COD: 138.6; SS: 358.7.
Hg: 23.8; Pb: 3.1; Cd: 0.4; Mg: 34.8; Ca: 49.2; Na:
Sample 2 291.4; P04 : 11.7; N03 : 16.4; NH4 : 28.7; Cl: 342.34;
BOD: 78.8; COD: 231.4; SS: 352.6.

These types of equations can be used to predict the Hg(II) removal for any
PGBS-COOH dose within the particular experimental conditions. To test the
robustness of the relationship presented in Equation (4.36) and (4.37) the values of
R (% Hg(II) removal) for different ms values were calculated from the above
equations. The comparison of the observed and predicted values of R for different
values of ms is shown in Figure 37. In each case the experimental and predicted
values are in excellent agreement.
168
Kinetic and Equilibrium Models ...........

120

-
100

� 80 Equilibrium time : 4 h
0
Temperature: 30 °c
60 pH :6.5

40 Initial Hg(II) concn.


__ : Theoretical Sample I : 50.0 mlefL
20
+o : Experimental Sample 2: 23.8 mlefL
0
0.0 1.0 2.0 3.0 4.0
Adsorbent dose (g/L)

Figure 37
Effect of adsorbent dose for the removal of Hg(II) from
chlor-alkali industry wastewater by PGBS-COOH

4.16.2 Testing with synthetic wastewater of Pb(II) and Cd(II)

The effectiveness of the adsorbent material has been confirmed by treating it with
synthetic wastewaters containing Pb(II) and Cd(II) ions in the presence of other
cations and anions. Composition of synthetic wastewater for Pb(II) and Cd(II) is
given in Table 14. The effect of adsorbent dose on Pb(II) and Cd(II) removal by
PGBS-COOH is given in Figure 38. The percentage removal of metals was found
to increase with increase in mass of adsorbent. This may be due to the increase in
adsorbent surface area with increase in mass of adsorbent. Almost complete
removal of Pb(II) from 50 mL wastewater containing 10.0 mg/L was possible with
100 mg (2.0 g/L) of the adsorbent. But for Cd(II) the complete removal from 50
mL wastewater containing 10.0 mg/L was possible with 160 mg (3.2 g/L) of
169
Chapter4

Table 14
Composition of synthetic wastewaters
containing Pb(II) and Cd(ll)
Metal Composition
ions (mg/L)

Pb(Il) Pb2+: 10; Ca2+: 60; Mg2+: 14.0; Na+: 20.0; K+: 3.0; so/-: 170.0; Cl 1-:
30.0; N03 1-: 3.0; Si02 : 8.0; (pH: 6.5)

Cd(Il) cct2+: IO; Mg2+: 5.0;1 Na+: 200.0;1


K+: 6.5; so/-·: 20.0; co/-: 5.0;
B40/-: 16.0; N03 -·: 380.0; F -: 45.0; (pH: 6.5)

+ +
PGBS-COOH. Considering the presence of other ions like, Ca2+, Mg2 , Na and
K+, the amounts of PGBS-COOH needed for the complete removal of Pb(II) and
Cd(II) f.,.om the wastewater, are in good agreement with that obtained from the
batch experiments discussed above. Examination ofFigure 38 also reveals that at a
fixed metal concentration, the percentage removal increases with increasing
adsorbent dose. However, unit adsorption or adsorption per unit mass ofadsorbent
decreases and the effect is more pronounced at higher adsorbent dose. Adsorption
of Pb(II) per unit weight of adsorbent decreases from 7.48 mg/g (59.8 %) at 0.40
g/L of adsorbent to 2.08 mg/g (100 %) at 2.40 g/L and for Cd(II), adsorption
decreases from 3.61 mg/g (28.9 %) at 0.40 g/L ofadsorbent to 1.56 mg/g (100 %)
at 3.20 g/L for the synthetic wastewater under consideration. This is because, as
the dose of adsorbent (g/L) is increased, there is less commensurate increase in
adsorption resulting from the lower utilization of the adsorption capacity of the
adsorbent. The results of the experiment were utilized to develop a mathematical
relationship between percentage removal (R) and adsorbent dose (ms) [using
boundary conditions] with r2 values greater than 0.99 and is represented as
170
Kinetic and Equilibrium Models ...........

R = ms
3 (4.37)
(3.47 x 10- )+(8. 75 x 10- 3) ms

(4.38)

The above equations corresponding to Pb(II) (4.37) and Cd(II) (4.38) can be used
to predict the Pb(II) and Cd(II) removal R (%) for any PGBS-COOH dose (ms)
within the particular experimental conditions. To test the validity of the
relationship presented in the above equations, the values of R (%) for different ms
values were calculated from the above equations. The comparison of the observed
and predicted values of R for different values of ms is shown in Figure 38. The
experimental and predicted values are in excellent agreement and suggest the
effectiveness of the equations.

120

100

'$. 80

,.· Pb Cd
] 60 Theoretical: -­
Experimental: <)
Initial concentration: 10 mglL

� 40 ,•
,-· Equilibrium time: 3 h
20 Temperature: 30 °c
pH:6.5
0
0.0 1� 2� 3� 4� 5�
Adsorbent dose (g/L)

Figure 38
Effect of adsorbent dose for the removal of Pb(II) and Cd(II)
from simulated wastewater by PGBS-COOH
171
Chapter4

4.16.3 Testing with simulated nuclear power plant coolant water


containing Co(II)
The suitability of the adsorbent material for treating Co(II) solutions was tested
using simulated nuclear power plant coolant water. The composition of water
samples is presented in Table 15 (Rengaraj and Moon, 2002). Figure 39 shows the
effect of the mass of PGBS-COOH onto the adsorption of Co(II) from coolant
waters. The removal of Co(II) was found to increase with increase in mass of
adsorbent. This may be due to the increase in adsorbent surface area with increase
of mass of adsorbent. It is evident that quantitative removal of Co(ll) ions from
SO mL synthetic water samples containing 1 and 10 mg/L metal ion and several
other ions a minimum adsorbent dose of 50 mg is sufficient for a removal of
81.8 and 56.8 % of Co(II) respectively. Almost complete removal of Co(II) from
SO mL �amples was possible with 75 and 250 mg of the adsorbent dose for Sample
I and 2 respectively which is in good agreement with that obtained from the batch
experiments mentioned above.

Table 15
Composition of Co(Il) containing synthetic nuclear
power plant coolant water
Concentration of element
Compound
Element
chosen (mg/L)
Sample I Sample 2
Sb(V) Sb20s 5 5
Co(Il) Co(N03)2.6H20 10
Fe(III) Fe(N03)3 .9H20 30 30
Ni(II) Ni(N03)2.6H20 15 15
Ag(l) AgN03 5 5
B(III) H3B03 20 20
Cr(III) Cr(N03 )3 .9H20 4 4
Li(I) LiOH.H20 0.5 0.5
Cs(I) CsN03 0.5 0.5
172
Kinetic and Equilibrium Models ...........

Removal of Cr(III) and Ni(II) from nuclear power plant coolant water by
adsorption on PGBS-COOH was also studied. About 98.8 % Cr(III) and 99. l %
Ni(II) from Sample 1 and 96.3 % Cr(III) and 97.1 % Ni(II) from sample 2 were
observed at pH 6.5 and an adsorbent dose of 100 mg in 50 mL. The results clearly
indicate that PGBS-COOH can also be used to remove Cr(III) and Ni(II) from
aqueous solutions. Further work is now under way to establish the effectiveness of
PGBS-COOH in removing other toxic heavy metals from wastewaters.

120

100

80
Equilibrium time : 4 h
pH :6.5
60

40 . Initial Co(II) concn.


Sample I: 10.0 mg'L
-- : Theoretical
20 Sample 2: 1.0 mg'L
• 0 : Experimental
0
0.0 2.0 4.0 6.0 8.0
Adsorbent dose (g/L)
Figure 39
Effect of adsorbent dose for the removal of Co(II) from
simulated nuclear power plant coolant water by PGBS-COOH

4.17 Desorption and Regeneration Studies

To make the adsorption process more economical it is necessary to regenerate the


spent adsorbent. Reusability of an adsorbent is of crucial importance in industrial
practice for metal removal from wastewater. Desorption studies help to elucidate
the mechanism of adsorption and recovery of the adsorbate and recycling of
adsorbent. Desorption of the adsorbed metal ions from the PGBS-COOH was
173
Chapter4

studied using a batch reactor system. The regeneration of the adsorbent may make
the treatment process economical.
The PGBS-COOH loaded with the maximum amount of M(II) ions was placed
in the desorption medium and the amount of metal ions desorbed in 4 h was
measured. Solutions having different concentrations of NaN03, Na2C03 , Na2S04,
NaCl, NaCl-HCl and HCl were evaluated for the extraction of adsorbed M(II) from
spent adsorbent. In this study, the ability of HCl to desorb the metal uptake by
PGBS-COOH was found to be maximum. From further experiments, 0.2 M HCl
was found to be a good reagent for the desorption of metal from the adsorbent.
Desorption efficiency is defined as the percentage extraction of metal initially
loaded onto the PGBS-COOH. Figure 40 clearly shows that the PGBS-COOH can
be used repeatedly without significantly losing its adsorption capacity. As shown in
this figure with four adsorption/desorption cycles, the adsorption (%) decreased
from 99.7, 99.3, 99.8 and 90.5 % to 91.7, 91.1, 90.2 and 79.21 % for Pb(II), Co(II),
Hg(Il) and Cd(II) respectively from Cycles 1 to 4. On the other hand, recovery of
M(II) ions in 0.2 M HCl decreased from 97.8, 94.8, 94.3 and 79.2 % in the first
cycle to 89.8, 86.5, 88.7 and 71.2 % in the fourth cycle respectively for Pb(II),
Co(II), Hg(II) and Cd(II). The reduction in metal uptake in four cycles of
adsorption-desorption indicates that the binding sites on the surface of the
adsorbent were destroyed or morphologically altered on repeated exposure to a
strong acid environment. The initial mass of adsorbent, 0.1 g used at the first cycle
was reduced to <0.085 g over four cycles of adsorption-desorption, resulting in an
adsorbent loss of more than 15 %. The loss of adsorbent might have contributed to
the reduction in metal uptake though it is not known whether HCI has the ability to
rupture the structure of the adsorbent. Great care had been taken to minimize the
loss of adsorbent and require further investigation.
174
Kinetic and Equilibrium Models ... ........

Figure 40 clearly shows that PGBS-COOH can be used repeatedly without


significantly loosing the adsorption capacity for metal ions. The ability of HCI at
0.2 M concentration to strip most of the adsorbed metal may be attributed to ion
+
exchange. It has been postulated that the high concentration of H ions at low pH is
responsible for the displacement of adsorbed Co(II) via ion exchange mechanism.
The small fraction of sorbed metals not recoverable by regeneration, presumably
represents the metal, which are bound through strong interaction, and, as a result,
the sorption capacity is reduced in successive cycles.

Pb(II) � : Adsorption Co(II) � : Adsorption


100
D: Desorption D: Desorption

90

=
?f!. 80

e,
.5!
70
Q

� 60
i
Q
100 Cd(II) � : Adsorption
D: Desorption
Hg(II) � : Adsorption
D: Desorption
Q

< 90

80

70

60 ....___.___.__..__..__..._....._...._........._._....._................__.___.__.�..__..._..._...._.........__.___.__._
1 2 3 4 2 3 4
Cycle number

Figure 40
Four cycles of adsorption/desorption of
Pb(II), Co(II), Hg(II) and Cd(II) onto/from PGBS-COOH
175
Chapter4

4.18 Batch Adsorption Design

The equilibrium relationship between the adsorbed metal amount per unit mass of
the adsorbent (qe) and the residual metal ion concentration {Ce) in solution may be
represented by isotherm models like Langmuir or Freundlich as discussed earlier.
In the present adsorption studies of lead(II), cobalt(II), mercury(II) and
cadmium(II) onto PGBS-COOH, the adsorption equilibrium was established in 3 h
and then the adsorbed amount did not change with time (Figure 21 ). The efficiency
of metal ion removal can be improved if the solution is treated with separate small
batches of adsorbent rather than in a single-stage batch process. Mutli-stage batch
process is important to maximize economical operation but filtration and handling
costs increase rapidly. Consequently two-stage batch adsorption process is
practically sound. Adsorption in batch stirred reactors in series can be considered
as a multistage equilibrium operation depending on any of the equation
corresponding to an isotherm model.
As seen in Figure 41, if the same volume of solution (Vo) is treated at each
stage by the same amount of adsorbent (m) to reduce the solute (metal ion)
concentration of the solution from Co to C/, the material balances for the metal ion
are for Stage l is

{4.39)

and for Stage 2

(4.40)

where C/ and C/, are the equilibrium concentration after the l st and 2nd stages
respectively and q/ and q/ are the amount of metal ions adsorbed per unit mass of
the adsorbent at the l st and 2nd stage respectively.
176
Kinetic and Equilibrium Models ...... .....

Equations (4.39) and (4.40) can be rearranged as for Stage 1:


I = m I
(Co -Ce) V (qe -qo) (4.41)
o
and for Stage 2:

(4.42)

The amount (q0) of metal adsorbed per unit mass of the adsorbent at the beginning
in each reactor is equal to 0.0. Equations (4.41) and (4.42) will provide the
operation lines passing through (q0, C0) and (qe 1, Ce 1 ) for the first stage and
(q0, Ce 1 ) and (q/, Ce2) for the second stage, respectively. Each slope is y = - m/V0,
i.e., the negative of the ratio of adsorbent quantity to wastewater volume. Since the
amounts of the adsorbent used in each stage are equal, the operation lines are
parallel.
m
Qo

Vo Vo Vo
I
Co

m
q/ q/

Figure 41
Schematic representation of a two-stage batch reactor
177
Chapter4

The equilibrium metal concentration of each stage can be found by using the
experimental equilibrium curve and operation line (Figure 42). To determine the
equilibrium metal ion concentration (Ce 1 ) in solution leaving the first reactor, the
initial metal ion concentration (Co) is located at the proper point on ordinate [q0=0,
C0 =300 mg/L each for Pb(II), Co(II), Hg(II) and C0 = 150 mg/L for Cd(II)]; then
the operation line passing through (qo, C0) with a slope of -m/V0 is drawn
(Figure 42). In the present case since 0.1 g of the adsorbent is added to 50 mL of
the solution the slope is -2.0. According to Equation (4.41), the intersection
coordinates of the operation line with the equilibrium curve give qe 1, c/ The
second stage is constructed by locating Ce I on the ordinate and drawing a straight
line (the second operation line) from (qo, Ce 1) with the same slope. The ordinate of
the intersection point of this line with the equilibrium curve gives the metal ion
concentration C/ at equilibrium in the effluent from the second reactor. A similar
construction for the third reactor leads to the desired concentration C/ leaving the
third reactor if the required concentration limit has not been attained. The residual
Pb(II), Co(II), Hg(II) and Cd(II) concentration leaving the second stages were
graphically found to be 15.2, 39.4, 58.5 and 15.0 mg/L respectively. The values of
q/ obtained from the operating lines corresponding to the second stage were found
to be 142.4, 130.3, 120.8 and 67.5 mg/g for Pb(II), Co(II), Hg(II) and Cd(II)
respectively at 30 °c. The qeI values of the metals corresponding to the first stage
of the metals were found to be closely agreeing with the qe values obtained from
the experiments and Langmuir equations.
The operational line thus helps in evaluating the theoretical number of stages
for the removal of M(II) ions from the wastewater. In the present study it was
found that M(II) would reduce virtually to 0.5 within three stages at 30 °c.
Additional research is warranted to examine the robustness of the above method
178
Kinetic and Equilibrium Models ...........

500 --����������--.500
Initial Pb(II) concn: 300 mg/L Initial Co(II) concn: 300 mg/L
Temperature: 30°C Temperature: 30°C
400 400

300 300

200 200

100 100

0 0
0 50 100 150 200 0 50 100 150
300 600
u Initial Cd(II) concn: 150 mg/L
Initial Hg(II) concn: 300 mg/L
°
250 Temperature: 30°C 500 Temperature: 30 C

200 400

150 300

100 200

50 100

0 0
0 15 30 45 60 75 0 50 100 150

Qe (mg/g)

Figure 42
Equilibrium curve and operation lines with mN0
slopes = -2.0 for Pb(II}, Co(II}, Hg(II) and Cd(II)
adsorption in two-stage batch reactors

with respect to real industrial wastewaters and other varying experimental


conditions.

You might also like