Sie sind auf Seite 1von 32

This article was downloaded by: [Indian Institute of Technology Roorkee]

On: 23 July 2014, At: 03:23


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK
Journal of Dispersion Science and Technology
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ldis20
Adsorptive Removal of Hg(II) From Synthetic and Real
Aqueous Solutions Using Modified Papaya Seed
Sunil Kumar Yadav
a
, Dhruv Kumar Singh
a
& Shishir Sinha
b
a
Analytical Research Laboratory, Department of Chemistry , Harcourt Butler Technological
Institute , Kanpur , U.P. , India
b
Department of Chemical Engineering , Indian Institute of Technology , Roorkee , U.K. ,
India
Accepted author version posted online: 16 Jun 2014.
To cite this article: Sunil Kumar Yadav , Dhruv Kumar Singh & Shishir Sinha (2014): Adsorptive Removal of Hg(II) From
Synthetic and Real Aqueous Solutions Using Modified Papaya Seed, Journal of Dispersion Science and Technology, DOI:
10.1080/01932691.2014.930713
To link to this article: http://dx.doi.org/10.1080/01932691.2014.930713
Disclaimer: This is a version of an unedited manuscript that has been accepted for publication. As a service
to authors and researchers we are providing this version of the accepted manuscript (AM). Copyediting,
typesetting, and review of the resulting proof will be undertaken on this manuscript before final publication of
the Version of Record (VoR). During production and pre-press, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal relate to this version also.
PLEASE SCROLL DOWN FOR ARTICLE
Taylor & Francis makes every effort to ensure the accuracy of all the information (the Content) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
1
Adsorptive Removal of Hg(II) from Synthetic and Real Aqueous Solutions Using
Modified Papaya Seed

Sunil Kumar Yadav
1
, Dhruv Kumar Singh
1
, Shishir Sinha
2

1
Analytical Research Laboratory, Department of Chemistry, Harcourt Butler
Technological Institute, Kanpur, U.P., India
2
Department of Chemical Engineering,
Indian Institute of Technology, Roorkee, U.K., India

Address correspondence to Dhruv Kumar Singh, Analytical Research Laboratory,
Department of Chemistry, Harcourt Butler Technological Institute, Kanpur 208002, U.P.,
India E-mail: dhruvks123@rediffmail.com, hbti_sky@rediff.com

Received 22 May 2014; accepted 31 May 2014.


Abstract
The performance of chemically modified papaya seed (CMPS) adsorbent with carboxyl
and amino groups has been studied. Adsorption experiments were performed with respect
to the changes in initial pH of the solution, contact time, initial Hg(II) concentration and
CMPS dosage. Kinetic data were fitted to the pseudo-second-order model. The maximum
adsorption capacity calculated by Langmuir model was 18.34 mg/ g. CMPS was
characterized by elemental analysis, Fourier transform infrared (FTIR) spectroscopy and
scanning electron microscopy (SEM). The results indicate that adsorption mechanism of
CMPS involves ion exchange (2Na
+
/Hg
2+
) and carboxylic group dominated surface
complexation. Regeneration study revealed that CMPS can be used successfully for four
cycles with a small adsorption capacity loss (6.8%).
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
2


KEYWORDS: Chemical modification papaya seed, mercury adsorption, kinetics,
isotherm

1. INTRODUCTION
An increasingly industrialized global economy over the last century has led to
many environmental problems. Heavy metals are common contaminants of industrial
wastewater.
[1,2]
Mercury is considered by the Environmental Protection Agency (EPA as
a highly dangerous element because of its accumulative and persistent character in the
environment and biota.
[3]
According to the Agency for Substances and Toxic Diseases
Registry (ATSDR), in 2011 mercury occupied the third position on the priority list of
hazardous substances.
[4]
The effluent discharged from chlorine chlor-alkali manufacturing
processes, pulp paper and oil refining, plastic, and battery manufacturing industries is an
important source of mercury pollution. The accumulation of mercury in environment,
which comes from municipal wastes and manufacturing of organo-mercurial compounds,
causes potential risk to human health due to the up taking of these amounts of mercury by
plants and their introduction into the food chain, including marine organisms (algae,
seaweed, fish, etc.).
[5]
Conventional methods for the removal of mercury from waste
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
3
streams include adsorption,
[611]
bio-sorption,
[12]
co-precipitation,
[13]
electrocogulation,
[14]

chemical precipitation,
[15]
membrane filtration,
[16]
solvent extraction,
[17]
reverse osmosis
and ion exchange.
[18]
However, most of these methods require either high energy or large
quantities of chemicals. Conventional precipitation methods do not always provide a
satisfactory removal rate to meet pollution control limits; moreover, synthetic ion-
exchange resins are often expensive, and adsorbents have low mechanical strength, weak
hydrothermal stability, or weak chemical bond with the metals.
[19]
It is important to
explore adsorptive material with low-cost, high adsorption speed, and good removal
performance for low concentration mercury(II). Recently, the synthesis of efficient and
eco-friendly adsorbents has attracted considerable attention.
[20,21,22]
Agro-waste, a kind of
low-cost and biodegradable adsorbent, is abundantly can be easily obtained in nature.
Low-cost sorbents reported to remove mercury with high efficiency includes lemna
minor,
[23]
palm shell,
[24]
terminalia catappa,
[25]
rubber waste,
[26]
mandarin peel,
[27]
and
Cassia javanica seed.
[28]

Attempts have been made to develop a low-cost and unconventional adsorbent
(CMPS) using papaya seed for the removal of Hg(II) from synthetic solution and chlor-
alkali wastewater. CMPS was characterized using elemental analysis, Fourier transform
infrared spectroscopy (FTIR) and scanning electron microscopy (SEM). The effects of
several operating parameters such as pH, initial Hg(II) ion concentration, contact time
and adsorbent dosage were investigated to arrive at optimum conditions for the
adsorption process. Various kinetic models as well as isotherm models have been studied
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
4
for their usefulness in correlating the experimental data. Reaction mechanism and
regeneration studies of CMPS for Hg(II) were also carried out.

2. MATERIALS AND METHODS
2.1 Reagents And Instruments
A stock solution (1000 mg/ L) of Hg(II) was prepared by dissolving 1.354 g of
mercury(II) chloride (E-Merck, Germany) in deionized water (DIW) and diluting up to 1
L. Ethylenediamine tetraacetic acid (EDTA), dimethyl sulfoxide (DMSO) and thionyl
chloride (SOCl
2
) were purchased from Merck, India. All other reagents were of AR
grade.

EI digital pH meter 111 (Chandigarh, India) was used for pH measurements.
Elemental analyses were performed using Perkin-Elmer 2400 Series CHNS Analyser,
Chennai (India). FTIR spectra were recorded in the frequency range of 4000-400 cm
-1
by
KBr pellet method using Bruker FTIR Vertex 70 (Germany). A rotatry shaking machine
(IEC-56) was used for shaking. The SEM micrographs of the adsorbents were obtained at
5.0 kV on a SUPRA 40 VP Field Emission Scanning Electron Microscopy (USA).
Systronic double beam spectrophotometer 2203, Ahmadabad, India, and Atomic
Absorption Spectrometer (AAS; model GBC 935) were used for determination of Hg(II).

2.2 Preparation Of CMPS
Biomass PS, collected from Kanpur (India) was washed with tap water followed
by DIW, sun dried for a week, powdered and sieved (125-105 m). The chemical
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
5
modification of PS was conducted in an anhydrous condition using DMSO as a solvent.
EDTA (11.68 g) and DMSO (100 mL) were mixed in a three-neck flask (250 mL). Then,
SOCl
2
(4.76 g) was slowly added to the mixture. Just after the addition of SOCl
2,
PS
(4.00 g) was quickly added to the mixture. The reaction was carried out at room
temperature under magnetic stirring for 2 h and chemically modified papaya seed was
obtained. Then, the CMPS was separated by vacuum filtration, washed in a row with
DMSO, acetone, DIW, sodium carbonate solution (Na
2
CO
3
, 0.1 mol/ L), DIW and
acetone. Finally, CMPS was dried at 802 C in a hot air oven, and kept in a desiccator
after cooling.

The preparation process of the CMPS is given in Scheme 1. In the carboxyl group
activation stage, the carboxyl group reacts with SOCl
2
, a reagent widely used in changing
the carboxyl into highly reactive acyl chloride.
[29]
The acyl chlorinated EDTA can be
grafted to CMPS matrix through acyl chloride group which reacts with hydroxyl and
amino groups on the surface of the PS and is grafted.

2.3 Adsorption Experiments
The adsorption experiments were carried out in Erlenmeyer flasks (150 mL)
containing 50 mL of Hg(II) solution (5-12 mg/ L). pH (1~6) was adjusted using NaOH or
HNO
3
before the addition of adsorbent and measured with pH meter. Then, the adsorbent
(0.025 g; particle size 125-105 m) was added and flasks were shaken in a rotator
shaking machine (160 rpm) at room temperature (30C) until the equilibrium was reached
(60 min). After equilibrium, the adsorbent was separated from the solution by filtration
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
6
using Whatman No.4 filter paper and the remaining concentration of Hg(II) ions in the
filtrate was analyzed by spectrophotometric method using dithizone (1,5-diphenyl
thiocarbazone) at 480 nm,
[30]
or AAS. The percentage of Hg(II) removal R and q
e

(mg/g) amount of Hg(II) adsorbed per unit mass of the adsorbent were determined using
the equations (1) and (2), respectively.
i e
i
C C
% R 100
C

=



(1)
i e
e
C C
q ( / ) V
M
mg g

=



(2)
where, C
i
and C
e
(mg / L) are the initial and equilibrium Hg(II) concentrations,
respectively, V is the solution volume (L) and M is the weight of the adsorbent (g).

3. RESULTS AND DISCUSSION
3.1Characterization Of PS And CMPS
3.1.1 Elemental Analysis
The chemical composition of adsorbents is important to explain its behavior
during sorption process. The results of C, H, N, S, and O analysis of the PS and CMPS
are presented in Table 1. The efficiency of grafting EDTA to the PS was evaluated by
comparing the oxygen and nitrogen contents before and after chemical modification.
Compared to PS, CMPS exhibits a sharp increase in oxygen (from 29.1 to 33.4%) and
nitrogen content (1.7 to 4.6%). Considering that no other carboxyl or nitrogen source
except EDTA is introduced during the whole preparation process, it could be concluded
that the EDTA is well grafted to PS.

D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
7
3.1.2 Sem Analysis
Scanning electron microscope is an extremely useful tool for visual confirmation
of surface morphology and the physical state of the surface. The SEM images of CMPS
and Hg(II)CMPS showed different surface characteristics. The surface of CMPS
appeared to be rough in nature, while some smoothness were observed on the surface
after Hg(II) adsorption, probably due to the formation of a mercury(II)CMPS complex,
involving carboxyl and amino functional groups (Figure 1a and b).

3.1.3 Ftir Analysis
FTIR spectra of PS, CMPS and Hg(II) loaded are shown in Figure 2 a and c. In
the FTIR spectra of PS and CMPS (Figure 2a, and b) the adsorption peak at 3375 and
3412 cm
-1
appear due to hydroxyl and amine groups. The peaks at ~2925 cm
-1
and 1656-
1646 cm
-1
may be attributed to alkyl CH
2
peaks and C=O of carboxyl groups/ ester/
amide, respectively. The peaks at wave number ~1244 and 1162 cm
-1
are due to C-N
stretching and /or C-O stretching vibrations. The absorption peaks in the range 1100-614
cm
-1
may be due to inorganic groups (e.g. phosphate and sulphur groups). The shifting of
the peaks from 3412; 1448 and 614 cm
-1
(Figure 2b) to 3376; 1384 and 627 cm
-1
(Figure
2c) seems due to adsorption of Hg(II) on CMPS.

3.2 Effect Of Solution Ph
The effect of pH on the adsorption of Hg(II) by CMPS at 30 C is shown in Figure
3. The results show that the percentage of removal increases from 4.3 to 63.7 with the
increase in pH (1 to 6). A slow adsorption occurred in the range of 1 < pH < 3; however,
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
8
a steep rise followed at pH > 3. The poor adsorption of Hg(II) ions at low solution pH
was due to the competitive coordination effect of the H
+
ions with N donor atoms of NH
and O donor atoms of OH on the surface of adsorbents.
[31]
Moreover, at lower pH, the
functional groups of CMPS might remain primarily in the protonated form, making it
difficult for Hg(II) adsorption. A more pronounced Hg(II) adsorption, in the pH range 3
5, was probably due to deprotonation of the functional groups on CMPS and interaction
of free Hg(II) ions with the active sites on CMPS. The uptake of Hg(II) at higher pH (>
5) is unlikely to be attributable to the formation of metal hydroxide species and/or
insoluble precipitates of Hg(OH)
2
. Thus, Hg(II) may interact with CMPS by a
complexation mechanism, involving the carboxyl, hydroxyl and amino functional groups,
and the maximum adsorption capacity (10.2 mg/ g) was achieved at pH 5. Consequently,
all the further experiments were performed at a pH of 5.

3.3 Effect Of Contact Time, Initial Concentration Of Hg(II) And Adsorbent Dosage
Figure 4(a) shows the effect of contact time on adsorption of Hg(II) ions by
chemically modified papaya seed. It has been observed that upto 30 min, there is sharp
increase in adsorption and then decrease. The decrease in adsorption after 30 min may be
due to decrease in binding sites of adsorbent and also due to decrease in Hg(II) ion
concentration in the solution. The rapid adsorption was desirable by virtue of the
economic viability.
[32, 33]
In the first 30 min, most of the active sites on the adsorbent
surface were vacant, and Hg(II) ion concentration in the solution was high. Along with
the adsorption processes going on, the amounts of available adsorption sites become less,
so only a very low increase in metal uptake was observed in the later stage.
[34]

D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
9

A higher initial Hg(II) concentration results in enhancing the adsorption process. The
equilibrium adsorption capacity of CMPS increases (6.5 to 12.5 mg/ g) or percentage
removal decreases (65 to 52.1%) with the increase in initial Hg(II) concentration (5-12
mg/ L). This explanation is in agreement with the experimental results (Figure 4b and c).
It could be at lower concentration, all Hg(II) ions in the solution would reach with the
binding sites and thus facilitated removal of almost by CMPS while at higher
concentration, more Hg(II) ions left unsorbed in the solution due to the saturation of the
binding sites. This indicates that the energetically less favorable sites become involves
with increasing Hg(II) ion concentration in aqueous solution.

The effect of CMPS dosage on adsorption of Hg(II) shows that the percentage of Hg(II)
removal from aqueous solution increases from 63.75 to 99.99% when adsorbent dosage
increases from 0.50 to 1.20 g/ L (Figure 4d). This may be due to an increased adsorbent
surface area and availability of more adsorption sites or more functional groups resulting
from the increased dosage of the adsorbent. After the critical dosage (1.2 g/ L), the
removal percentage remains unchanged.

3.4 Kinetics Of Adsorption
This study describes three kinetic models pseudo first order, pseudo second order
and intra-particle diffusion to predict the adsorption mechanism.

3.4.1 Pseudo First Order Kinetic Model
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
10
The pseudo first order model is given by:
[35]

( ) ( )
1
e t e
k
log q -q logq t
2.303

=


(3)
where, q
t
and q
e
are the amounts of Hg(II) adsorbed (mg/ g) at time t and at equilibrium,
respectively, and k
1
is the rate constant of pseudo-first-order adsorption (min
-1
). The
linearized form of the pseudo-first order model for the adsorption of Hg(II) ion onto
CMPS is given in Figure 5a. The values of the rate constant k
1
and adsorption capacity q
e

were obtained from the slopes and intercepts of the plots of log (q
e
- q
t
) versus t,
respectively (Table 2). A low correlation coefficient (R
2
) for the pseudo first order (0.96)
kinetic model suggested that the adsorption of Hg(II) deviated considerably from the
pseudo first-order kinetic model and may not be enough to explain the adsorption
mechanism of Hg(II) on the surface of CMPS.

3.4.2 Pseudo Second Order Kinetic Model
Ho and Mckay developed a pseudo second order equation based on the amount of
adsorbed adsorbate on the adsorbent
[36]
. It can be written as
2
t 2 e e
t 1 1
t
q k q q

= +


(4)
where, q
e
and q
t
are the adsorption capacities at equilibrium and time t (mg/ g),
respectively, k
2
is the rate constant of the pseudo-second order sorption (g/ mg

min). The
plots of t/q
t
versus t gave linear plot (Figure 5b). The values of the adsorption parameters,
q
e
and k
2
were determined from the slope and intercept of the plot, respectively (Table 2).
The value of the correlation coefficients (R
2
) for pseudo second-order kinetic model
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
11
obtained was 0.999 which indicated that the adsorption system studied belongs to the
second order kinetic model.

3.4.2 Intra-Particle Diffusion Model
The intraparticle diffusion model is expressed as
[37]

q
t
= K
id
t
0.5
+ C (5)
where, k
id
is the intra-particle diffusion rate constant (mg/ g min

), C is the intercept (mg/


g). The plot of q
t
versus t
0.5
for Hg(II) adsorption onto CMPS (Figure 5c) does not pass
through the origin. This indicates that intraparticle diffusion is not the only rate-limiting
step. There were three processes controlling the adsorption rate but only one was rate
limiting in any particular time interval. The values of correlation coefficients fitted with
kinetic models are presented in Table 2. The fitted linear regression plot showed that the
experimental data were well fitted to the pseudo-second-order model with better value of
correlation coefficient (R
2
= 0.999), compared to the pseudo-first-order kinetic and intra-
particle models. The q
e
,
cal
values were closer to the q
e,
exp values for pseudo-second-
order model.

3.5 Adsorption Equilibrium
Different isotherm models have been used to determine the adsorption efficiency
for adsorption of Hg(II) from aqueous solution. In this study, three adsorption isotherm
models (Langmuir, Freundlich and Temkin isotherms) were used for Hg(II) adsorption on
CMPS.

D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
12
3.5.1 Langmuir Isotherm Model
To ensure equilibrium conditions, the linear form of the Langmuir isotherm model
(Equation 6) was applied to the experimental data as:
[38]

e e
e M M
C C 1
q b Q Q
= + (6)
where, C
e
, q
m
and b are the concentration of adsorbate at equilibrium (mg / L), maximum
adsorption capacity (mg/ g) and Langmuir constant (L/ mg), respectively. The essential
characteristics of the Langmuir isotherm can also be expressed in terms of a
dimensionless constant separation factor or equilibrium parameter (R
L
), which is defined
by equation (7).
i
1
L 1 b C
R
+
= (7)
where, C
i
is the initial solute concentration and b is the Langmuirs adsorption constant
(L/ mg). The R
L
value confirms the adsorption to be unfavorable (R
L
> 1), linear (R
L
=
1), favorable (0 < R
L
< 1) or irreversible (R
L
= 0).
[39]
The value of R
L
for the adsorption
of mercury ions on CMPS was 0.012 to 0.004; which indicates that adsorption process is
favorable.

3.5.2 Freundlich Isotherm Model
The empirical Freundlich isotherm
[40]
is given by equation (8):
1
e F e n
lnq lnK lnC = + (8)
where, q
e
is the amount adsorbed (mg/ g) and C
e
is the equilibrium concentration of the
adsorbate (mg / L). The K
F
(constant related to adsorption capacity) and 1/n (constant
related with the adsorption intensity) are Freundlich constants. The value of n is an
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
13
indication of the favorability of adsorption. Values of n > 1 represent favorable nature of
adsorption.



The Langmuir and Freundlich adsorption isotherms of raw-PS are compared (Figure not
shown), adsorption data are given in Table 3. On the whole, the Langmuir isotherm
displays a higher regression coefficient (R
2
) compared to the Freundlich isotherm for
both raw-PS and CMPS and the adsorption data fit better with the Langmuir adsorption
isotherm model.

3.5.3 Temkin Isotherm Model
The Temkin isotherm model assumes that the adsorption energy decreases linearly with
the surface coverage due to adsorbentadsorbate interactions. The linear form of Temkin
isotherm model
[41]
is described by equation:
e e
q ln A lnC B B = + (9)
where, B = RT/b, b is the Temkin constant (J/ mol) related to adsorption heat, T is the
absolute temperature (K), R is the gas constant (8.314 J/ mol K), and A is the Temkin
isotherm constant (L/ g). B and A can be calculated from the slope and intercept of the
plot of q
e
versus ln C
e,
respectively (Figure 6c). The Temkin constants B, b, and A
together with the R
2
values are shown in Table 4. The heat of adsorption of Hg(II) in the
layer would decrease linearly with coverage due to adsorbate/adsorbate interactions.

Among all these three isotherms, Langmuir model (Figure 6a) fitted the experimental
data best (R
2
> 0.99). All of the R
L
values less than 1 for studied initial concentrations (5
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
14
to 12 mg / L) confirmed a favorable adsorption. The maximum monolayer Hg(II)
adsorption capacity for CMPS was determined as 18.34 mg /g. The value of n >1,
indicated favorable adsorption of Hg(II) onto CMPS.

3.6 Mechanism
The Hg(II) adsorption depends upon nature of the adsorbent surface. The results of FTIR
and elemental analysis demonstrated that the adsorption of Hg(II) by CMPS occurred by
a chemical adsorption mechanism, which means that N and O atoms of the functional
groups in CMPS seemed to complex with Hg(II) in aqueous solutions (Scheme 1). The
shifting of bands from 3412; 1448 and 614 cm
-1
(Figure 2b) to 3376; 1384 and 627 cm
-
1
(Figure 2c) seems due to adsorption of Hg(II) on CMPS through complexation. The
other possibility may be ion exchange (Hg
2+
/ Na
+
). Hence, adsorption mechanism of
Hg(II) on CMPS involves complexation as well as ion exchange.

3.7 Regeneration Of CMPS
In addition to good adsorption capacity, it is also desirable that an adsorbent can be
regenerated and reused repeatly with regard to the cost. The regeneration capacity of
CMPS was evaluated by HCl, H
2
SO
4
and HNO
3
acid treatment. 0.5 M HCl solution was
found more effective as an eluting reagent than other concentrations as well as other acid
types. 25 mg of Hg(II) adsorbed CMPS was treated with 50 mL of 0.5 M HCl solution,
under stirring for 2 h at room temperature. After acid regeneration and neutralization
(with a 0.1 M NaHCO
3
), adsorbent was for reused in three successive adsorption
regeneration cycles. The adsorption capacity of CMPS after fourth cycle was 9.2 mg/ g,
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
15
indicated a loss in adsorption capacity 6.8% compared to the initial capacity (10.2 mg/g).
This study reveals the economic cyclic use of the adsorbent.

3.8 Removal Of Hg(II) From Chloro Alkali Waste Water
In order to study the adsorption potential of the prepared adsorbent (CMPS) for the
removal of Hg(II) from real wastewater, batch experiments were conducted on chloro
alkali wastewater. Characteristics of chloroalkali waste water are given in Table 4. As it
is evident from the data, chloro-alkali wastewater contains a high concentration of
different ions, e.g. Cl
-
, Na
+
, Ca
++
and Mg
++
, which can interfere during the adsorption of
Hg(II) on the CMPS. Therefore, the selectivity of adsorbent is an important factor. CMPS
is a good adsorbent with high capability of Hg(II) ion removal (99.2%) from chloro-alkali
wastewater.

4. CONCLUSIONS
Carboxylic and amino functionalized papaya seed was successfully prepared for the
removal of Hg(II) ions from synthetic and chloro-alkali wastewater. The FTIR, SEM and
elemental analysis confirmed the linking of functional groups on CMPS. The adsorption
experiments revealed that the optimum pH is 5 for maximum removal of Hg(II) ions. The
maximum adsorption capacity of CMPS for Hg(II) ions removal was found at 18.2 mg/ g
by fitting with Langmuir isotherms model. The adsorption desorption batch experiments
showed a small change in adsorption capacity of 6.8% for Hg(II) after four cycles,
demonstrating a good regeneration capacity of CMPS for wastewater treatment
containing Hg(II) ions. Finally, we observed that the developed CMPS is a good
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
16
adsorbent for removal of Hg(II) ions by 99.2 % from chloro-alkali wastewater. CMPS
can be synthesized through easy route using economic adsorbent and its disposal is
environmentally safe.

ACKNOWLEDGMENTS
The authors are thankful to Director, HBTI Kanpur for providing necessary research
facilities.

REFERENCES
[1] Gupta, V. K., Agarwal, S., and Saleh, T. A. (2011) Water Res., 45(6): 2207-12.
[2] Gupta, V. K., and Nayak, A. (2012) Chem. Eng. J., 180: 81-90.
[3] Karunasagar, D., Krishna, M. V. B., Rao, C. V., and Arunachalam, J. (2005) J.
Hazard. Mater., B118: 133-139.
[4] ATSDR. Detailed data for 2011 priority list of hazardous substances [Internet].
ATSDR, Division of Toxicology and Environmental Medicine [cited Sept 2013].
Available from: http://www.atsdr.cdc.gov/spl/
[5] Sanchez Uria, J. E., and Sanz-Medel, A. (1998) Talanta, 47: 509-524.
[6] Liang, S., Guo, X., and Tian, Q. (2013) Desalin. Water Treat., 5: 7166-7171.
[7] Yadav, S. K., Singh, D. K., and Sinha, S. (2014) J. Environ. Chem. Eng., 2: 9-19.
[8] Kosak, A., Lobnik, A., and Bauman, M. (2013) Int. J. Appl. Ceram. Technol., 1-
12.
[9] Gupta, V. K., Carrott, P. J. M., Carrott, M. M. L., and Suhas (2010) Critical Rev.
Environ. Sci. Technol., 39: 783-842.
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
17
[10] Gupta, V. K., Rastogi, A., and Nayak, A. (2010) J. Colloid Inter. Sci., 342: 783-
842.
[11] Gupta, V. K., Pathania, D., Sharma, S., Agarwal, S., and Singh, P. (2013) J.
Molecu. Liq., 177: 343-352.
[12] Doke, K., Khan, E., and Gaikwad, V. (2013) Journal of Dispersion Science and
Technology, 34: 1347-1355.
[13] Lv, X., Jiang, G., Xue, X., Wu, D., Sun, T. S., and Xu, C. X. (2013) J. Hazard.
Mater., 262(15): 748-758.
[14] Xu, H., Yang, Z., Ming Zeng, G., Luo, Y., Huang, J., Wang, L., Song, P., and
Mo, X. (2014) Chem. Eng. J., 239 (1): 132-140.
[15] Oncel, M. S., Muhcu, A., Demirbas, E., and Kobya, M. (2013) J. Environ. Chem.
Eng., 1(4): 989-995.
[16] Zhou, L., Zhang, Z., Jiang, W., Guo, W., Ngo, H., Meng, X., Fan, J., Zhao, J.,
and Xia, S. (2014) Biofouling: The Journal of Bioadhesion and Biofilm Research, 30:
105-114.
[17] Luo, J. H., Li, J., Qi, Y. B., Cao, Y. Q. (2013) Desal. Water Treat., 51: 2130-
2134.
[18] Ali, S. W., Malik, M. A., Yasin, T., Muhammad, B. (2013) J. Appl. Polym. Sci.,
129 (4): 2234-2243.
[19] Wu, X. W., Ma, H. W., Li, J. H., Zhang, J., Li, Z. H. (2007) J. Colloid Interface
Sci., 315: 555-561.
[20] Mittal, A., Mittal, J., Malviya, A., and Gupta, V. K. (2009) J. Colloid Interface
Sci., 340 (1): 16-26.
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
18
[21] Gupta, V. K., and Ali, I. (2007) Water Res., 41 (15): 3307-3316.
[22] Mittal, A., and Gupta, V. K. (2008) J. Hazard. Mater., 151 (2-3): 821-832.
[23] Xing, L. S., Ying, Z. F., Yang, H., and Cong, N. J. (2011) J. Hazard. Mater., 186:
423-429.
[24] Kushwaha, S., Sodaye, S., Padmaja, P. (2008) World Acad. Sci. Eng. Technol.,
43: 600-606.
[25] Inbaraj, B. S., and Sulochana, N. (2006) J. Hazard. Mater., B133: 283-290.
[26] Vizuete, E. M., Garcia, A. M., Gisbert, A. N., Gonzalez, C. F., and Serrano, V. G.
(2005) J. Hazard. Mater., B119: 231-238.
[27] Dalal. Z. H. (2013) Desal. Water Treat., 51: 6761-6769.
[28] Singh, V., Singh, S. K., and Maurya, S. (2010) Chem. Eng. J., 160: 129-137.
[29] Li, C. S., Tsai, Y. H., Lee, W. C., and Kuo, W. J. (2010) J. Org. Chem., 75:
4004-4013.
[30] Khan, H., Ahmed, M. J., and Bhanger, M. I. (2005) Anal Sci., 21 (5): 507-12.
[31] Liu, C. K., Bai, R. B., and Ly, Q. S. (2008) Water Res.,42: 1511-1522.
[32] Mittal, A., Mittal, J., Malviya, A., and Gupta, V. K. (2010) J. Colloid Interface
Sci., 344 (2): 497-507.
[33] Gupta, V., Gupta, B., Rastogi, A., and Agarwal, S. (2011) J. Hazard. Mater., 186
(1): 891-901.
[34] Gupta, V. K., and Rastogi, A. (2010) J. Colloid Interface Sci., 342 (2): 533-539.
[35] Lagergren, S. (1898) Handlingar, 24 (04): 1-39.
[36] Ho, Y. S., and McKay, G. (2000) Water Res., 34: 735-742.
D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
19
[37] Weber, W. J., and Morris, J. C. (1963) J. Sanitary Eng. Div., Am. Soc. Civil Eng.,
89: 39-59.
[38] El Qada, E. N., Allen, S. J., and Walker, G. M. (2006) Chem. Eng. J., 124: 103.
[39] Nasir, M. H., Nadeem, R., Akhtar, K., Hanif, M. A., and Khalid, A. M. (2007)
J. Hazard. Mater., 147: 1006-1014.
[40] Freundlich, H. (1907) Z. Phys. Chem., 57: 385-470.
[41] Mane, V. S., Mall, I. D., and Srivastava, V. C. (2007) J. Environ. Manage., 84:
390-400.















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
20
Table 1 Physicochemical characteristics of CMPS
Parameters Value
pH 6.7
Bulk density (g cm
-3
) 0.63
Moisture content (%) 11.7
Solubility in water (%) 0.0
Solubility in 0.5 M HCl (%) 0.0
surface area (m
2
/

g) 19.48
Element analysis C% H% N% S% O
PS 52.9 7.2 1.7 2.1 36.1
CMPS 43.7 9.1 4.6 2.4 40.2












D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
21
Table 2 Pseudo-first-order, pseudo-second order and intra particle diffusion rate for the
adsorption of Hg(II)

by CMPS
Constants Values
q
e exp.
(mg/ g) 10.2
Pseudo-first order
q
e (cal)
(mg/ g) 6.85
k
1
(min
-1
) 0.021
R
2
0.961
Pseudo-second order
q
e (cal)
(mg/ g) 10.9
k
2
(g/ mg

min) 0.011
R
2
0.999
Intraparticle diffusion
k
id
(mg/ g min) 0.639
R
2
0.815








D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
22
Table 3 Parameters of Langmuir, Freundlich and Temkin adsorption isotherm constants
for the adsorption of Hg(II) by CMPS and raw-PS
Isotherm Constants Values
CMPS Raw-PS
Langmuir Q
m
(mg/ g) 18.34 10.75
b (L/ mg) 0.38 0.112
R
2
0.998 0.993
R
L
0.012- 0.005 0.018- 0.007
Freundlich K
F
(mg/ g) 5.12 2.013
n 1.85 5.02
R
2
0.917 0.977
Temkin B 4.14 -
b(J/ mol) 39.87 -
A (L/ g) 3.37 -
R
2
0.954 -







D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
23
Table 4 Characteristics of chloro-alkali wastewater
Parameters Value
pH ~7.8
Hg(II) 54.8g/ L
Pb(II) <0.02 mg / L
Cd(II) <0.01 mg / L
Mg(II) 85 mg / L
Ca(II) 195 mg / L
Na(I) 305 mg / L
Cl
-
1081 mg / L
NO
3
-
<0.05 mg / L
Conductivity 2700 s/ cm
Turbidity >1000 NTU








D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
24
Scheme 1. Synthetic route to CMPS.





















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
25
Scheme 2. Proposed structure of binding mechanism of Hg(II) on to CMPS.




















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
26
Figure 1. SEM of (a) CMPS (b) Hg(II) loaded CMPS.

















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
27
Figure 2. FTIR spectra of (a) CMPS (b) Hg(II)-loaded CMPS.















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
28
Figure 3. Effect of pH for the adsorption of Hg(II) by CMPS (Initial concentration = 8
mg/ L, adsorbent dose = 0.5 g/ L, contact time = 60 min).

















D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
29
Figure 4. (a) Effect of contact time for the adsorption of Hg(II) by CMPS (Initial pH = 5;
initial concentration = 8 mg/ L, adsorbent dose = 0.5 g/ L). (b and c) Effect of initial
concentration for the adsorption of Hg(II) by CMPS (Initial pH= 5, contact time = 60
min, adsorbent dose = 0.5 g/ L). (d) Effect of adsorbent dosage on Hg(II) removal (Initial
pH= 5, initial concentration= 8, mg L
-1
, contact time: 60 min).








D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
30
Figure 5. (a) Pseudo-first order plot (b) Pseudo-second order plot and (c) Plot of intra
particle diffusion model for adsorption of Hg(II) by CMPS (Initial pH = 5, adsorbent
dose = 0.5 g/ L, contact time = 60 min).












D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
31
Figure 6. (a) Langmuir isotherm (b) Freundlich isotherm and (c) Temkin isotherm plots
for adsorption of Hg(II) by CMPS (Initial pH= 5, adsorbent dose = 0.5 g /L, contact time
= 60 min).

D
o
w
n
l
o
a
d
e
d

b
y

[
I
n
d
i
a
n

I
n
s
t
i
t
u
t
e

o
f

T
e
c
h
n
o
l
o
g
y

R
o
o
r
k
e
e
]

a
t

0
3
:
2
3

2
3

J
u
l
y

2
0
1
4

Das könnte Ihnen auch gefallen