Вы находитесь на странице: 1из 11

Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

Contents lists available at SciVerse ScienceDirect

Journal of the Taiwan Institute of Chemical Engineers


journal homepage: www.elsevier.com/locate/jtice

An investigation into the adsorption removal of ammonium by salt activated


Chinese (Hulaodu) natural zeolite: Kinetics, isotherms, and thermodynamics
Aref Alshameri a, Chunjie Yan a,*, Yasir Al-Ani b, Ammar Salman Dawood b, Abdullateef Ibrahim c,
Chunyu Zhou a, Hongquan Wang a
a
b
c

Engineering Research Center of Nano-geomaterial of Education Ministry, China University of Geosciences, Wuhan 430074, China
School of Environmental Studies, China University of Geosciences, Wuhan 430074, China
State Key Laboratory of Biogeology and Environmental Geology, China University of Geosciences, Wuhan 430074, China

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 2 February 2013
Received in revised form 2 May 2013
Accepted 6 May 2013
Available online 12 June 2013

The development of the process of sodium activation of zeolite has been an effective technique for
enhancing the efciency of ammonium removal. In this research, the optimum conditions for the
activation of Chinese (Hulaodu) zeolite of the most effective parameters such as sodium concentration,
stirring time, and temperature were determined. The most efcient conditions were selected according
to the highest ammonium removal capacity. The characteristics of activated zeolite (ActZ) and its
mechanism of ammonium removal were investigated and compared with that of natural zeolite (NZ).
Additionally, both zeolites were analyzed by scanning electron microscopy (SEM), Zeta potential, X-ray
diffraction (XRD), thermogravimetry (TG) and BET surface analysis. The activated zeolite revealed the
highest ammonium removal efciency reaching up to 98% based on stirring time, zeolite loading, initial
ammonium concentration, temperature and pH. The adsorption kinetic was explored and tted best
with the pseudo-second-order model, whereas adsorption isotherm results illustrated that Langmuir
model (LM) provided the best t for the equilibrium data. Moreover, thermodynamic parameters such as
change in free energy (DG8), enthalpy (DH8) and entropy (DS8) were also calculated. The parameters
revealed that the exchange of ammonium ion by activated zeolite occurred spontaneously at ambient
conditions (25 8C). It was concluded that when Chinese (Hulaodu) zeolite is activated under the
condition of 1 M NaCl, 70 8C and stirring time of 30 min, an excellent removal of NH4+ was obtained.
2013 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords:
Zeolite (clinoptilolite)
Adsorption
Activated zeolite
Ammonium removal
Kinetics and isotherm

1. Introduction
Clinoptilolite (Na,K,Ca)6(Si,Al)36O7220H2O is one of the naturally existing zeolites. It is rich in silica and has a lower ionexchange capacity than other zeolites as well as less than that of
many of the available synthetic ion-exchange resins. It generally
exhibits a high selectivity for ammonium and metallic ions [1,2].
Clinoptilolite from different deposits has been widely reported as
adsorbent for ammonium removal from wastewaters [3,4]. Natural
zeolite, on the other hand, needs to be puried and modied in
order to improve its ion-exchange [5] and adsorption properties
before it can be used to remove ammonium effectively. Nitrogen
compounds in aqueous environments are usually found in the form
of ammonium ions (NH4+). Important sources of NH4+ include
efuent from municipal sewage treatment plants, the application
of fertilizer in agricultural practices and industrial processes all of

* Corresponding author. Tel.: +86 18971579917; fax: +86 027 67885098.


E-mail address: chjyan2005@126.com (C. Yan).

which contributes to the accelerated incidence of eutrophication


resulting in algal bloom in lakes and rivers [610]. Complete
removal of NH4+ is required due to its toxicity to the majority of
aquatic life. For example, the ammonium nitrogen concentration
for most sh species must not exceed 1.5 mg NH4+ ion [11,12].
NH4+ concentration, in certain surface waters serving as a source of
potable water, is much higher than the permissible level, due to
large quantities of industrial and municipal wastewater being
discharged into existing water resources [5,13]. This threatens the
availability of safe drinking water and, thus, human health. For this
reason, the prevention of nitrogen pollution with NH4+ removal
from wastewater is of great importance [7,9,14,15].
Various methods including air stripping, biological methods
and activated carbon have been used for NH4+ removal [1619].
However, since biological methods do not respond well to shock
loads of ammonium, unacceptable peaks of NH4+ over the
discharging levels may frequently appear in the efuent. Also,
high costs, poor regeneration and uncertainty of outcome are some
of the frequently encountered limitations in the application of the
biological method. Moreover, there is a high risk to safety during

1876-1070/$ see front matter 2013 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jtice.2013.05.008

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

Nomenclature

555

2. Materials and methods


2.1. Raw materials

the starting equilibrium concentrations (mg/L)


C0
the nal equilibrium concentrations (mg/L)
Ce
the volume of the working solution (L)
V
the mass of added zeolite (g)
M
k1 and k2 (1/min) and (g/mg min) are constants of
adsorption
the rate of adsorption (mg/g min)
h
Langmuir constant (L/mg)
K
Freundlich adsorbent capacity (mg/g (L/mg)1/n)
KF
the reciprocal of reaction order
n
adsorption capacity at time t (mg/g)
qt
adsorption capacity at equilibrium conditions
qe
(mg/g)
maximum adsorption capacity (mg/g)
qmax
constant of intraparticle diffusion (mg/g min1/2)
Kid
a and k0 constants of Bangham equation
the volume of solution (ml) of Bangham equation
V
the weight of adsorbent (g/L) of Bangham equation
m

the subsequent processing such as the aeration process which


causes stripping effects for volatile compounds resulting in
accidental releases, often causing odor and aerosol with health
implications [20,21].
Contingency on temperature and climate conditions constitutes
another disadvantage in this process [22]. Compared with the
above mentioned methods, high safety, low cost [13,2326] and
relative simplicity of application and operation are some of the
attributes that are attracting an increasing focus on the use of
zeolite for environment applications [2,8,11].
The factors that inuence ammonium removal performances
are mainly pH, temperature, reaction time, initial concentration
of NH4+, and adsorbent dosage. Although many previous studies
have focused on these factors collectively [3,27,28], however,
comparison of the results from the available literature indicates
signicant variability in the reported behavior.
It seems that natural clinoptilolites from different places have
different characteristics [9,10,29,30]. These differences in the
characteristics of clinoptilolites are probably attributed to the
differences in the geological formation of zeolite sources
[21,31,32]. Therefore, each special zeolite material has its own
special characteristics and still requires to be researched individually [11,27,31].
The main aim of this study is to determine the optimal
conditions for the activation of Chinese (Hulaodu) zeolite to get the
best results for the adsorption of NH4+ by salt treatment. The main
focus of the study is the transformation of low grade Hulaodus
zeolite to a high cation exchanger under appropriate activation
conditions. An investigation and comparison of characteristics and
the equilibrium removal of NH4+ ion onto both zeolites was also
carried out.
The specic objectives of this research is to study the sodium
activation of zeolite samples and the effect of various parameters
on zeolite activation such as sodium concentration, stirring time,
and temperature. And to identify the key processes controlling the
rate of ammonium adsorption by zeolite. In addition, the effects of
pH, stirring times, initial concentration, adsorbent dosages and
temperature on NH4+ removal for both natural and activated
samples were investigated and compared. Adsorption isotherms,
thermodynamics and reusability of zeolite for the removal of NH4+
ions were examined.

A natural zeolite (NZ) was collected from Huludao city in China.


Analytical grade inorganic chemicals, such as ammonium chloride
(NH4Cl), sodium chloride (NaCl), sodium hydroxide (NaOH) and
hydrochloric acid (HCl), were used.
2.2. Preparation and activation of zeolite
A natural zeolite (NZ) was ground and passed through 200
230 mesh sieves. The material was washed with distilled water to
remove any non-adhesive impurities, and then dried in an oven at
100 8C for 24 h and nely crushed. The activation process was
carried out by mixing NZ powder material with an aqueous
solution of sodium chloride under the different conditions detailed
below:
Effect of sodium concentration: To study the effect of sodium
concentration and batch activation, the sodium chloride
concentration was varied from 0.5 to 2 mol/L with zeolite/
solution ratio maintained as 1 g/10 ml. The suspension was
stirred in conical asks (500 ml) using a magnetic stirrer
water bath at a rate of 120 rpm and 90 8C for 2 h. Subsequently,
the suspension was ltered and washed with distilled water.
The wet activated material was dried at 70 8C in an oven for 24 h
and used in batch adsorption experiments at an initial
ammonium concentration of 100 mg/L with a pH 7 at 25 8C
and stirring time of 24 h. The best sodium ion concentration for
the activation of zeolite was selected as that corresponding to
the highest NH4+ removal capacity (mg/g).
Effect of time: For determining the optimum time required for
activation zeolite from aqueous solution, a weighed quantity of
zeolite (1 g) was added into the solution of 1 mol/L sodium
chloride concentration and stirring time ranging from 0.5 to 3 h.
The conditions applied above on the effect of sodium
concentration experiment were repeated. The optimum stirring
time for the activation of zeolite was selected as that
corresponding to the highest NH4+ removal capacity (mg/g).
Effect of temperature: The effect of temperature on activation of
zeolite was investigated at different temperature values
ranging from 10 to 90 8C. The sodium chloride concentration
and stirring time were kept constant at 1 mol/L and 0.5 h,
respectively. The optimum temperature for the activation of
zeolite was selected as that corresponding to the highest NH4+
removal capacity (mg/g).
2.3. Analysis and characterization
The activated and natural zeolites were characterized by XRD,
SEM, EDX, TG, Zeta potential, chemical analysis and specic surface
area (BET). Identication of mineral species in the zeolite samples
was carried out by XRD pattern using a Germany D8-FOCOS X-ray
diffractometer with Cu Ka (l = 0.154 nm) radiation operating at
40 kV and 40 mA and a step width of 0.058. Semi-quantitative
weight percentages of samples were calculated by using mineral
intensity factors. Textural characteristics of the activated and
natural zeolites were performed using a Japanese Netherlands
FESEM Quanta SU8010 electron microscope, operating at an
accelerating voltage of 15 kV for photomicrographs as well as to
analyze the Chinese-zeolite composition by Energy Dispersion Xray Spectrometry (EDS), USA, Apolloxp. The sample was initially
placed in a vacuum chamber for coating with a thin layer (a few
nanometers) of gold (Au). The specic surface area, pore size and
volume of the material were evaluated by the nitrogen gas

556

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

adsorption method, with a heating rate of 10 8C/min. N2


adsorptiondesorption experiments were performed at 77 K with
an Automatic Volumetric Sorption Analyzer (ASAP2020, TSI,
America), employing multipoint BET isotherm adsorption data
tting. Zeta potential measurements for the natural and activated
zeolites, as a function of medium pH, were determined using Zeta
Plusk equipment (zetasizer Nano ZS 90, Malvern, UK). A 103 mol/L
solution of KNO3 and 60 mg/L of ammonium concentration were
used. The medium pH was controlled by adding HCl (pH < 7) and
NaOH (pH > 7) solutions, separately. The water content and zeolite
decomposed were determined by a Thermogravimetric (TG)
analyzer. TG/DSC analysis was performed on a NEZSCH, STA409PC-Germany, and thermal analysis system was in the range 30
1200 8C. In addition, the composition of zeolite was analyzed by
chemical method. Absorbance values of ammonium ion concentration in solution were read using a Chinese-Shimadzu UV-723Vis
spectrophotometer. Quality control testing includes experiments
with blanks and duplicates.
2.4. Adsorption rate, batch sorption studies
All adsorptions in batch experiments were carried out using
stopper conical asks (500 ml), zeolite/liquid ratio of 1 g/100 ml,
magnetic stirring water bath, a stirring rate of 120 rpm and a
temperature at 25 8C. A stock solution (1000 mg/L) was prepared
by dissolving NH4Cl in distilled water.
For determining the optimum time required for ammonium
removal by natural zeolite (NZ) and activated zeolite (ActZ) from
aqueous solution, a weighed quantity of adsorbent (1 g) was added
into solution of 80 mg/L ammonium concentration and was stirred
for 10420 min and 560 min of NZ and ActZ, respectively, at a xed
pH of 7. The effect of initial ammonium concentration in batch
adsorption experiments was carried out by using initial ammonium
concentration in the range of 10240 mg/L and 10400 mg/L for
300 min and 40 min for NZ and ActZ, respectively. The effect of pH on
adsorption was investigated at initial ammonium concentration of
80 mg/L and performed in different pH values (210) at 25 8C. Batch
adsorption was conducted at 300 and 40 min for NZ and ActZ,
respectively. The pH of the solution was adjusted by 1 M HCl or
NaOH solution. For determining zeolites loading effect, zeolite
loading was varied from 0.2 to 2.2 g/100 ml at initial ammonium
concentration of 80 mg/L. The zeoliteliquid was then stirred for 300
and 40 min for NZ and ActZ, respectively, at temperature 25 8C.
Temperature of adsorption isotherms was studied at 25, 35 and
45 8C. A 100 ml of ammonium chloride solution of 80 mg/L
concentration was equilibrated with 1 g zeolite for 40 min for
ActZ. Samples were ltered through a 0.45 mm lter membrane
after adsorption.
The residual concentration of ammonium was determined by
Nesslers reagent spectrophotometry method at 420 nm.
The removal efciency (%) of zeolite and the amount of
adsorbed ammonium ions (qe) were calculated, respectively, using
the following equations:
Removal efficiency %

qe

C0  Ce
 100
C0

C0  Ce
M

(1)

(2)

3. Results and discussion


3.1. Activated zeolite
Due to different results in activation of different zeolite
samples, parameters, such as sodium concentration, stirring time

and temperature, had to be investigated to obtain optimum


activation conditions, and the results are shown in Fig. 1ac. As can
be seen from batch adsorption experiments (Fig. 1ac), the
temperature of 70 8C, stirring time of 30 min and 1 mol/L
concentration of NaCl were the most effective in adsorption of
NH4+ ions on Hulaodu natural zeolite (NZ). The optimum
conditions obtained above for the activation of zeolite using NaCl
were applied on NaOH for the adsorption of NH4+ ions at 100 mg/L
of ammonium concentration, pH 7 and 25 8C. Its NH4+ adsorption
capacity (mg/g) was compared with NaCl activation as shown in
Table 1. The results showed that NaCl-activated zeolite (ActZ) had
a higher NH4+ adsorption capacity value than NaOH-activated
zeolite. This result was conrmed by EDS analysis (Fig. 2a and b)
which showed that the Na+ content of the NaOH-activated zeolite
(1.33%) was lower than that of NaCl-activated zeolite (2.10%).
Higher NH4+ adsorption capacity means higher Na+ content
[3234]. From the above discussion, it can be concluded that ion
exchange promoted by NaCl-activated zeolite (ActZ) is a more
attractive zeolite preparation method than that of NaOH-activated
zeolite. Therefore, NaCl-activated zeolite (ActZ) was selected for
further study.
3.2. Natural and activated zeolite characteristics
SEM and the qualitative composition analysis of EDS obtained
from the grain of natural and activated zeolites are shown in Fig. 2.
Fig. 2c illustrates that the main chemical elements (Al, Si, O, Na, Mg,
K, Ca and Fe) are present in the structure of this natural zeolite, in
agreement with the chemical composition (Table 2). Quantitative
tests were also performed on both zeolites. The results from
elemental analysis by EDS of NZ and ActZ as shown in Fig. 2a and c
indicate that the Na+ content in NZ increased from 0.21% to 2.10%
after activation. Meanwhile, the content of Ca2+, Mg2+ and K+
decreased from 3.44%, 0.81% and 1.53% to 1.08%, 0.58% and 0.76%,
respectively. For this reason, the ammonium capacity of combined
activated zeolite increased sharply in microscale [33,34].
The XRD patterns of NZ and ActZ are shown in Fig. 3a. X-ray
diffraction did not show any changes in zeolite structure after
activation. Mineralogical analysis of the zeolite samples was
carried out using X-Ray Diffraction (XRD). The results showed that
the natural zeolite contained clinoptilolite in the majority 93%, and
small quantity of quartz 7%.
Chemical components of natural zeolite are shown in Table 2.
The Si/Al ratio calculated from these data was found to be 4.8. The
ratio of Si/Al is an important factor in understanding the zeolite
structure. When the ratio is over than 4.0, then the zeolite is a
clinoptilolite-type and as such, the structure would not be broken
easily at high temperature.
This result was conrmed by thermogravimetric (TG) analysis
as shown in Fig. 3b. It shows that the zeolite could stand
temperatures of up to 869.2 8C without being decomposed and that
it only undergoes a weight reduction of 0.07% at this temperature.
It also, shows that the water content of zeolite was lost at 71.2 8C.
The surface analysis of activated zeolite was investigated by
Zeta potential and BET standard method and compared with that of
natural zeolite. Fig. 3c shows that zeolite surface groups are mainly
negative in the studied pH range and that the ammonium removal
does not interfere much with the Zeta potential measurements,
conrming the theory that the mechanism is not electrostatic
(charge neutralization) but a result of an ion-exchange reaction.
Moreover, the gure shows that activated zeolite is more negative
than natural zeolite.
The BET specic surface area, total pore volume, and average
width pore size of the natural zeolite were measured to be
25.88 m2/g, 0.0032 cm3/g and 8.72264 mm, respectively. However, these values did not change signicantly after activation of

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

Fig. 1. Removal capacity as a function of sodium concentrations (a), stirring time (b)
and temperature (c).

zeolite, 26.7074 m2/g of specic surface area, 0.004167 cm3/g of


total pore volume and 9.64864 mm of average width pore size.
These characteristics of activated zeolite have disadvantages for
physical adsorption, but the results from equilibrium experiments do not match with this fact because the removal
mechanism of ammonium by zeolites follows the ion-exchange
reaction. Therefore, ion-exchange capacities of ammonium
depend on dielectric strength between ammonium as well as

557

Fig. 2. SEM and EDS spectra analyses zeolite grain of NaCl-activated zeolite (ActZ)
(a), NaOH-activated zeolite (b) and natural zeolite (NZ) (c).

their afnities to zeolite. The size of micropores of the natural


zeolite is in the range of 310 A and micropores of the NaClmodied zeolite were more developed at the range of 5 A than
other cases. This implies that the NaClzeolite could selectively
remove NH4+ ions which have a specic size. Also, owing to its
small surface area, this zeolite is very stable against heat and
acidity as reported by Tehrani and Salari [29] and Gottardi and
Galli [35] who found that heat and acid activation does not result

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

558

Table 1
Comparison of NaCl and NaOH in zeolite activation at 100 mg/L of ammonium
concentration, pH 7, 25 8C and stirring time of 24 h.
Sodium ion used for activation Sodium concentration Ammonium removal
(mol/L)
capacity (mg/g)
NaCl-activated zeolite
NaOH-activated zeolite

1
1

5.921
4.552

in drastic changes in zeolite structure but increases Si/Al ratio, as


a consequence of heat or acid activation.
3.3. Effect of stirring time
As shown in Fig. 4a and b, the removal efciency of NH4+ ions
increased with increasing stirring time. 40% and 92% of NH4+ ions
removal were completed within 10 min for NZ and ActZ,
respectively, which also conrmed larger adsorption capacity of
ActZ compared to NZ.
The experimental data show that NZ could exceed 78%
percentage removal at 300 min, but then, the removal efciency
plateaus. As to ActZ, the ammonium removal rate exceeds 98%
percentage removal at 40 min and became increasingly slow with
increasing stirring time. This may be attributed to the utilization of
the most readily available adsorption sites of the zeolite that leads
to a fast diffusion and rapid equilibrium attainment .On the basis of
these results, 300 min and 40 min stirring period was selected for
all further studies of NZ and ActZ, respectively. Beyond this level
there is no noticeable increase in the adsorption and it is thus xed
as the equilibrium time. It can be said that the NH4+ ions were
adsorbed by the exterior surface of the adsorbent. When the
adsorption of exterior surface of the adsorbent reached the
saturation point, the NH4+ ions enter the adsorbent pores and
are adsorbed by the interior surface of the particles [34,36,37].
3.4. Effect of initial ammonium concentration
As shown in Fig. 5, the increment of removal efciency was
achieved in the ranges of 1050 mg/L of NH4+ concentrations for
both NZ and ActZ. This result is similar with that reported by
Sarioglu [11] who concluded that the increase in removal
efciency was achieved between 8.8 and 40 mg/L of ammonium
concentrations, indicating that the initial NH4+ concentration plays
an important role in the adsorption of ammonium onto zeolites.
Increasing the initial NH4+ concentration would increase the
mass transfer driving force and therefore the rate at which
ammonium ions pass from the bulk solution to the particle surface
[37]. The result can be generally expected for clinoptilolite having
micropores and macropores [7]. After 50 mg/L NH4+ concentration
of both zeolites, the removal efciency of ammonium decreased
with increased initial NH4+ concentration. This is because the high
initial ammonium concentration provided a greater driving force
[38]. As a result, the NH4+ ion could migrate from the external
surface to the internal micropores of the zeolite within a given
stirring time [32]. The equilibrium was achieved when all the
exchangeable ammonium and cation on the external and internal
surfaces of the zeolite were reached [31]. Fig. 5 shows that the
adsorption capacity of the activated zeolite was higher than that of
natural zeolite at each initial ammonium concentration.

Fig. 3. The XRD of activated and natural zeolites (a), TG of natural zeolite (b) and
Zeta potential of activated and natural zeolites (c).

3.5. Effect of solution pH and the mechanism of adsorption


The pH of the aqueous solution is an important factor
controlling ammonium adsorption [21]. As shown in Fig. 6, pH
played an important role for NH4+ adsorption of NZ and ActZ. The
removal efciency of both zeolites increased with increasing pH

Table 2
Chemical composition of the natural zeolite (wt.%).
NZ

SiO2

Al2O3

TFe2O3

TiO2

MgO

Na2O

CaO

K2O

P2O5

MnO

H2O

L.O.I.

66.34

12.23

0.99

0.16

0.98

0.73

3.17

1.37

0.027

0.026

5.06

13.88

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

559

Fig. 4. Effect of stirring time on NH4+ removal capacity of natural zeolite (a) and activated zeolite (b) (initial NH4+ concentration: 80 mg/L, 25 8C and pH 7).

from 2 to 7 and then it decreased gradually from pH 8 to 10 with


the maximum value being achieved at pH 7. An almost similar
trend has been reported for ammonium adsorption onto zeolite by
different researchers [21,30]. The behavior of ammonium as a
function of waters pH can be explained by considering the change
in density of hydrogen ions, the dominant ionic species of
ammonium and the surface charge of zeolite as a function of
waters pH. The pHzpc of natural zeolite and activated zeolite were
7.7 and 7.8, respectively, implying that the zeolite particles surface
is uncharged at water pH of 7.7 and 7.8 of natural and activated
zeolites; the zeolites particles surface has a positive charge at
water pH below 7.7 and 7.8, and it is negatively charged at water
pH above 7.7 and 7.8. This indicates that the surface of zeolite is
negatively charged at pH of above 7 [39]. For ammonium
adsorption, according to the relation of ammonium dissociation
in water as a function of pH, NH4+ ions is the dominant species of
ammonia nitrogen in water at pH of below 7 while the molecular
form, ammonia (NH3) is dominant at alkaline pH. By considering
the above facts, the increase in adsorption of ammonium with the
increase in waters pH up to the maximum point (pH 7) can be
attributed to a decrease of hydrogen ions in solution corresponding
to an increase in pH, and thus a reduction of the competition
between hydrogen ions and NH4+ ions for adsorption/exchanging
sites onto zeolite particles [6,40]. The decrease of ammonium
adsorption with the increase of waters pH above 7 is related to the
increase in percentage of molecular ammonium, which resulted in
the reduction of ionexchange potential [10].

Fig. 5. Effect of initial ammonium concentration on NH4+ removal capacity of the


ActZ and NZ (stirring time: 40 min and 300 min for ActZ and NZ, respectively, at
25 8C and pH 7).

3.6. Effect of adsorbent dosage


As illustrated in Fig. 7, the removal efciency of NH4+ ions by
both zeolites increases with increasing amount of both zeolites.
This effect can be attributed to an increased surface area and
number of adsorption sites [41]. As can be seen in Fig. 7, the
ammonium removal rate of ActZ increases more rapidly than that
of NZ and attains a plateau at 98.73% when the adsorbent dosage
was 1 g indicating that the NH4+ ions removal was negligible at
higher adsorbent dosage. The natural zeolite attains a plateau at
94.5% when the dosage was 1.8 g as shown in Fig. 7. Thus, both of
them reached a balance of approximately 98% when the adsorbent
dosage for ActZ and NZ was 1 g and 1.8 g, respectively. This may be
attributed to a large adsorbent amount which effectively reduces
the unsaturation of the adsorption sites and correspondingly, the
number of such sites per unit mass comes down resulting in
comparatively less adsorption at higher adsorbent amount [37,41].
Hence, when the ammonium is exchanged completely with cations
on the zeolite surface at a certain amount of zeolite loading, the
NH4+ removal reached equilibrium.
3.7. Kinetics of ammonium exchange
To identify the key process controlling the adsorption rate,
several models must be checked for suitability and consistency over
a broad range of the system parameters. Pseudo-rst- and -secondorder and Bangham equations as well as the diffusion-based

Fig. 6. Effect of pH on the removal of NH4+ ions (adsorbent dosage: 1 g/100 ml;
stirring time: 40 min and 300 min for ActZ and NZ, respectively; initial NH4+
concentration: 80 mg/L at 25 8C).

560

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

Fig. 7. Effect of adsorbent dosage on the removal of NH4+ ion (stirring time: 40 and
300 min for ActZ and NZ, respectively; initial NH4+ concentration: 80 mg/L at 25 8C
and pH 7).

WeberMorris model are used to t the experimental data and can


be summarized as follows.
The linear form of pseudo-rst, pseudo-second order models for
boundary conditions of q = 0 at t = 0 and qt = qe at t = te are as
follows:
The pseudo-first-order Eq: : lnqe  qt ln qe  k1 t

(3)

t
1
1

t
The pseudo-second-order Eq: :
qt k2 q2e qe

(4)

h k2 q2e





C0
k0 m
log
a log t
Bangham Eq: : log log
230V
C 0  qt m

(5)

The adsorption kinetics of ammonium by ActZ and NZ are


presented in Fig. 8ae. The kinetic data were better tted by the
pseudo second order model than other models as indicated by
higher R2 values (Table 3). Also it shows a higher sorption rate for
ActZ than NZ.
The pseudo second order model indicates that chemisorption
dominated in the adsorption process [42]. The difference in the
adsorbed concentration of adsorbate at equilibrium (qe) and at
time t (qt) is the key driving force for the adsorption, and the
adsorption capacity is proportional to the number of active
adsorption sites on the adsorbent [43]. There are three steps
involved in pseudo second order kinetic model: (i) the ammonium
ions diffuse from liquid phase to liquidsolid interface; (ii) the
ammonium ions move from liquidsolid interface to solid
surfaces; and (iii) the ammonium ions diffuse into the particle
pores [23]. Herein, the diffusion of ammonium ions from aqueous
phase was much faster than the surface and intraparticle diffusion
processes because the adsorption was performed under stirring
conditions [22].
To reveal the relative contribution of surface and intraparticle
diffusion to the kinetic process, the kinetic adsorption data were
further tted with the WeberMorris model using Eq. (6).
WeberMorriss Eq: : qt kid t 1=2 C

(6)

Intraparticle diffusion is assumed to be the sole rate-controlling


step if the regression of qt versus t1/2 is linear and the plot passes
through the origin [44]. Our tting results show that the regression
was linearly, but the plot did not pass through the origin (C 6 0).
Therefore, the adsorption kinetics of NH4+ ions on zeolite was
regulated by both surface and intraparticle diffusion processes. As

can be seen from Fig. 8d and e, ammonium exchange by both


zeolites involves two stages. These two stages suggest that the
ammonium exchange process proceeds by surface sorption and
intraparticle diffusion. It has been suggested that the rst one can
be attributed to the instantaneous occupation of most available
surface sites by exchanging NH4+ ions onto zeolites particles. The
surface of zeolites was negatively charged at water pH below 7
(Fig. 6) and its rate is very fast. The second region is due to a gradual
adsorption stage, where the ammonium ions enter into zeolites
particle by intraparticle diffusion through the pores. The values of
intercept C provide information about the thickness of the
boundary layer and the resistance to the external mass transfer
increases as the intercept increases. The constant C was found to
increase from NZ (1.3642 mg/g) to ActZ (3.7406 mg/g) as shown in
Table 3, which indicates the increase of the thickness of the
boundary layer and decrease of the chance of the external mass
transfer and hence increase of the chance of internal mass transfer
[37,41].
Table 3 presents the results of tting experimental data to the
pseudo-rst, pseudo-second-order, Bangham and intraparticle
diffusion models. It can be seen from Table 3 that the correlation
coefcient R2 varies in the order: pseudo-second order >
Bangham > intraparticle diffusion > pseudo-rst order model
under all experimental conditions, which indicates that the
pseudo-second-order model is most suitable in describing the
adsorption kinetics of ammonium on zeolite.
3.8. Ammonium exchange isotherms
Isotherm tting with model equations is a key issue to explore
adsorption mechanisms. Langmuir model (LM) and Freundlich
model (FM) were evaluated as follows:
LM is based on the assumption that each active site can only
hold one adsorbate molecule. The linear form of LM is expressed
as:
Ce
1
1

Ce
kqmax qmax
q

(7)

FM endorses the heterogeneity of the surface and assumes that


the adsorption occurs at sites with different energy of adsorption.
It is described as:
log q log kF

1
log C e
n

(8)

The linear plot of Langmuir isotherm of ActZ and NZ is shown in


Fig. 9a. It is noted that the values of qmax and k were calculated from
the slope and the intercept of the plot using Eq. (7) and are given in
Table 4. It will be seen that applicability of the simple Langmuir
equation for the present isotherm data indicates that the Langmuir
equation was able to properly describe the isotherm of ammonium
on the two zeolites (correlations coefcient >0.97). As shown in
Table 4, the ActZ had much higher ion exchange capacity than NZ. A
comparison with other zeolites from various literature reviews,
Nguyen and Tanner [14] reported a qmax value of 5.76 mg/g
ammonium adsorption on Australian natural zeolite. It has also
been reported that the qmax of ammonium removal using a Turkish
natural clinoptilolite was 8.1 mg/g [8]. Meanwhile a qmax value of
0.085 mg/g was reported by Demir et al. [36].
Fig. 9b shows the linearized Freundlich adsorption isotherm of
ammonium curve and the Freundlich parameters is presented in
Table 4. The FM of ActZ provides a slightly more consistent t to the
data compared with the FM of NZ. Similar values of 1/n which are
less than 1 have been reported for NH4+ removal using natural
zeolites from different countries [11,36].
To sum up, for the two zeolites, the experimental data are better
tted by the LM than FM as can be seen from the higher R2 in

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

561

Fig. 8. Pseudo-rst order (a), pseudo-second-order (b), Bangham kinetic plots (c) and intra-particle diffusion of both zeolites ActZ (d) and NZ (e) for NH4+ removal.

Table 4. In Langmuir and Freundlich equations, coefcients K and


KF represent the maximum amount that can be sorbed. Both values
for K and KF indicated that the activated zeolite has the higher
sorption capacity than the natural zeolite as shown in Table 4.
Table 3
Kinetic parameters for NH4+ removal using various kinetic models.
Kinetic model

Parameters

Pseudo-rst order

K1 (min1)

qe (mg/g)

R2

Absorbent
ActZ
NZ

0.0888
0.0165

0.119123
2.5126

0.954
0.918

Kinetic model

Parameters

Pseudo-second order

h (mg/g min)

qe (mg/g)

R2

Absorbent
ActZ
NZ

2.7156
0.1646

3.821
3.28

0.999
0.997

Kinetic model

Parameters

Bangham model

K0

R2

Absorbent
ActZ
NZ

201.609
25.616

0.0379
0.2981

0.979
0.958

Kinetic model

Parameters

Intraparticle diffusion model

Kid

R2

Absorbent
ActZ
NZ

0.0102
0.0977

3.7406
1.3642

0.957
0.966

3.9. Thermodynamic study


The effect of temperature on ammonium exchange was studied
at 25, 35 and 45 8C. Fig. 10 indicates that the amount of ammonium
exchanged onto zeolite increases with a decrease in temperature. A
similar trend was also observed of some adsorbents, including
Turkish clinoptilolite [8], Turkish zeolite [3], yash and sepiolite
[40], and NaA zeolite [45]. This may be due to a tendency for the
ammonium molecules to escape from the solid phase to the bulk
phase with the solution temperature increase [3,45]. In contrast,
increases in the NH4+ adsorption capacity with increasing
temperature have been reported for some other adsorbents
[2,9]. Therefore, by comparing the results of the present work
with those of the literature, it can be concluded that the effect of
temperature on the adsorption of ammonium depends on both the
nature of the adsorbent and the selected experimental conditions
[21]. Furthermore, ammonium exchange capacity decreases with
increasing temperature due to a weakening of the attractive forces
between NH4+ and adsorbent sites [41] and when the temperature
increases, solubility of ammonium increases and its adsorption
decreases [37]. The thermodynamic parameters, such as Gibbs
energy (DG8), enthalpy (DH8), and entropy (DS8), for the adsorption
of ammonium on zeolites were calculated using the following
equations:
K0

qe
Ce

DG RT ln K 0

(9)
(10)

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

562

Fig. 9. The linearized Langmuir (a), Freundlich (b) of ActZ and NZ and Langmuir (c), Freundlich (d) adsorption isotherm of NH4+ curve at different temperature of ActZ.
Table 5
Change of thermodynamic parameters with temperature.

DG DH  T DS


ln K 0 

DH 
RT

(11)

DS

(12)

Eq. (12) represents a mathematical relationship between K0 and


1/T. The values of K0, DG8, DH8 and DS8 parameters are summarized
in Table 5. Change in the standard free energy DG8 has negative
values (2.8662 and 1.3052 kJ/mol) at 25 and 35 8C, respectively,
but positive value (0.224 kJ/mol) at 45 8C. The negative values of
free energy change (DG8) indicate that this adsorption process is
spontaneous; therefore, no energy input to the system is required.
The higher negative value reects a more energetically
favorable adsorption [46]. For that reason, more energetically
favorable adsorption occurs at 25 8C. Change in the standard
enthalpy DH8 indicates a negative value of 49.384 kJ/mol;
therefore, ammonium exchange is an exothermic process. Also,

Temperature (8C)

K0

DG8 (kJ/mol)

DH8 (kJ/mol)

DS8 (kJ/mol)

25
35
45

2.89
2.01
0.82

2.8662
1.3052
0.224

49.384

0.1561

the negative value of the standard entropy change DS8 (0.1561 kJ/
mol) suggests that the randomness decreases the removal of NH4+
ions on the clinoptilolite [47].
The linear plot of LM and FM isotherm of ActZ at 25, 35 and
45 8C is shown in Fig. 9c and d. The LM and FM parameters are
presented in Table 5. The maximum value of K and KF at 25 8C
indicates that the NH4+ adsorption process is most effective at
25 8C. Comparing the correlation coefcients in Table 5 reects

Table 4
Constants for equilibrium isotherm models of NZ and ActZ.
Isotherm model

Parameters

Langmuir

Qmax (mg/g)

K (L/mg)

R2

9.515
9.533
9.794

0.444
0.313
0.123

0.9982
0.9986
0.967

3.445

0.1998

0.9772

ActZ
25 8C
35 8C
45 8C
NZ
25 8C
Isotherm model

Parameters

Freundlich

KF ((mg/g)/(mg/L)1/n)

1/n

R2

3.5743
3.0634
2.0305

0.2992
0.3043
0.3938

0.9716
0.9442
0.8867

1.252

0.5192

0.7363

ActZ
25 8C
35 8C
45 8C
NZ
25 8C

Fig. 10. Effect temperature on the exchange isotherm of NH4+ on activated zeolite
(ActZ).

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

References

Table 6
Data about adsorption and regeneration of zeolite.
Cycles

Ammonium adsorption
capacity (mg/g)

563

5.633

5.03

4.51

3.78

3.11

that LM yields a much better (R2 = 0.9670.998) t than that of the


FM (R2 = 0.8870.971).
3.10. Desorption and reusability
Accordingly, the ion-exchange and electrostatic adsorptions are
probably the main mechanisms for NH4+ ions removal by zeolite
under the selected conditions, through the following reactions
(Eq. (13)):
Ammoniumremoval : zeoliteNa NH4
! zeoliteNH4 Na

(13)

In this study, adsorption experiments were performed using 1 g


zeolite and 100 ml of 60 mg/L NH4+ at 25 8C for 40 min, and
desorption of adsorbed NH4+ onto zeolite was studied using 1 mol/
L NaCl, with zeolite/liquid ratio of 1 g/10 ml for 30 min at 70 8C,
and consecutive adsorptiondesorption cycles were repeated four
times. The results are shown in Table 6. There was a slight decrease
with the increase of cycle times in adsorption efciency from
5.633 mg/g for the rst cycle to 3.11 mg/g for the fth cycle. The
zeolite that was regenerated by four cycles in sodium solution had
high ammonium-removal efciency with its adsorption efciency
of 3.11 mg/g which is close to that of natural zeolite (3.344 mg/g).
4. Conclusion
The characteristics of Chinese (Hulaodu) natural zeolite (NZ),
activated zeolite (ActZ) and their efcacy in removing ammonium
were investigated. The following conclusions were made from the
experimental results:
 1 mol/L of NaCl, stirring time of 30 min and 70 8C were found to
be the optimum activation conditions for the zeolite with
excellent removal of NH4+.
 The highest adsorption capacity was obtained at pH 7 for both
zeolites and the maximum ammonium adsorption was rapidly
attained within 40 min and 300 min for activated zeolite and
natural zeolite, respectively.
 Langmuir adsorption isotherm of both zeolites t well with the
equilibrium adsorption data, and this adsorption process agrees
very well with pseudo-second-order kinetics rate model.
 DG8, DH8 and DS8 values reveal the exothermic and spontaneous
nature of the process and low temperature (25 8C) favors the
NH4+ removal on the zeolite.
 The adsorption capacity of activated zeolite is decreased to
3.11 mg/g which is close to that of natural zeolite (3.344 mg/g)
after being regenerated four times. Based on these results, the
study shows that Chinese (Hulaodu) natural zeolite can be used
as cheap, efcient and ecofriendly adsorbent for removing
ammonium from water and wastewaters.
Acknowledgment
This study was supported by Engineering Research Center of
Nano-Geomaterial of Education Ministry, China University of
Geosciences, Wuhan.

[1] Shavandi MA, Haddadian Z, Ismail MHS, Abdullah N, Abidin ZZ. Removal of Fe
(III), Mn (II) and Zn (II) from palm oil mill efuent (POME) by natural zeolite. J
Taiwan Inst Chem Eng 2012;43:7509.
[2] Du Q, Liu S, Cao Z, Wang Y. Ammonia removal from aqueous solution using
natural Chinese clinoptilolite. Sep Purif Technol 2005;44:22934.
[3] Saltali K, Sari A, Aydin M. Removal of ammonium ion from aqueous solution by
natural Turkish (Yildizeli) zeolite for environmental quality. J Hazard Mater
2007;141:25863.
[4] Arslan A, Veli S. Zeolite 13X for adsorption of ammonium ions from aqueous
solutions and hen slaughterhouse wastewaters. J Taiwan Inst Chem Eng
2012;43:3938.
[5] Li M, Zhu X, Zhu F, Ren G, Cao G, Song L. Application of modied zeolite for
ammonium removal from drinking water. Desalination 2011;271:295300.
[6] Huang H, Xiao X, Yan B, Yang L. Ammonium removal from aqueous solutions
by using natural Chinese (Chende) zeolite as adsorbent. J Taiwan Inst Chem
Eng 2010;175:24752.
[7] Jorgensen TC, Weatherley LR. Ammonia removal from wastewater by ion
exchange in the presence of organic contaminants. Water Res 2003;37:17238.
[8] Karadag D, Koc Y, Turan M, Armagan B. Removal of ammonium ion from
aqueous solution using natural Turkish clinoptilolite. J Hazard Mater 2006;136:
6049.
a A, Berber-Mendoza MS,
[9] Leyva-Ramos R, Monsivais-Rocha JE, Aragon-Pin
Guerrero-Coronado RM, Alonso-Davila P, et al. Removal of ammonium from
aqueous solution by ion exchange on natural and modied chabazite. J Environ
Manage 2010;91:26628.
[10] Wang S, Zhu ZH. Characterization and environmental application of an Australian natural zeolite for basic dye removal from aqueous solution. J Hazard
Mater 2006;136:94652.
[11] Sarioglu M. Removal of ammonium from municipal wastewater using natural
Turkish (Dogantepe) zeolite. Sep Purif Technol 2005;41:111.
[12] Barry FJ. The evolution of the enforcement provisions of the federal water
pollution control act: a study of the difculty in developing effective legislation. Mich L Rev 1970;68:110330.
[13] Dimirkou A, Doula MK. Use of clinoptilolite and an Fe-overexchanged clinoptilolite in Zn2+ and Mn2+ removal from drinking water. Desalination
2008;224:28092.
[14] Nguyen ML, Tanner CC. Ammonium removal from wastewaters using natural
New Zealand zeolites. New Zeal J Agric Res 1998;41:42746.
[15] Gupta VK, Rastogi A, Nayak A. Biosorption of nickel onto treated alga (Oedogonium hatei): application of isotherm and kinetic models. J Colloid Interface
Sci 2010;342:5339.
[16] Gupta VK, Gupta B, Rastogi A, Agarwal S, Nayak A. A comparative investigation
on adsorption performances of mesoporous activated carbon prepared from
waste rubber tire and activated carbon for a hazardous azo dyeAcid Blue 113.
J Hazard Mater 2011;186:891901.
[17] Mittal A, Mittal J, Malviya A, Kaur D, Gupta VK. Adsorption of hazardous dye
crystal violet from wastewater by waste materials. J Colloid Interface Sci
2010;343:46373.
[18] Moradi O, Yari M, Zare K, Mirza B, Naja F. Carbon nanotubes: a review of
chemistry principles and reactions. Fuller Nanotub Carbon Nanostruct
2012;20:13851.
[19] Moradi O, Zare K. Adsorption of ammonium ion by multi-walled carbon
nanotube: kinetics and thermodynamic studies. Fuller Nanotub Carbon
Nanostruct 2013;21:44959.
[20] European Commission. IPPC reference document on best available techniques
in common waste water and waste gas treatment/management systems in the
chemical sector. Sevilla: European IPPC Bureau; 2003.
[21] Moussavi G, Talebi S, Farrokhi M, Sabouti RM. The investigation of mechanism,
kinetic and isotherm of ammonia and humic acid co-adsorption onto natural
zeolite. Chem Eng J 2011;171:115969.
[22] Liao P, Yuan S, Xie W, Zhang W, Tong M, Wang K. Adsorption of nitrogenheterocyclic compounds on bamboo charcoal: kinetics, thermodynamics and
microwave regeneration. J Colloid Interface Sci 2013;390:18995.
[23] Liao P, Ismael ZM, Zhang W, Yuan S, Tong M, Wang K, et al. Adsorption of dyes
from aqueous solutions by microwave modied bamboo charcoal. Chem Eng J
2012;195:33946.
[24] Alshameri A, Abood AR, Yan C, Muhammad AM. Characteristics, modication
and environmental application of Yemens natural bentonite. Arab J Geosci
2013;113.
[25] Gupta VK, Jain R, Nayak A, Agarwal S, Shrivastava M. Removal of the hazardous
dye-tartrazine by photodegradation on titanium dioxide surface. Mater Sci
Eng C 2011;31:10627.
[26] Jain AK, Gupta VK, Bhatnagar A, Suhas. A comparative study of adsorbents
prepared from industrial wastes for removal of dyes. Sep Sci Technol 2003;38:
46381.
[27] Wang YF, Lin F, Pang WQ. Ammonium exchange in aqueous solution using
Chinese natural clinoptilolite and modied zeolite. J Hazard Mater 2007;142:
1604.
[28] Gupta VK, Jain R, Mittal A, Saleh TA, Nayak A, Agarwal S, et al. Photo-catalytic
degradation of toxic dye amaranth on TiO2/UV in aqueous suspensions. Mater
Sci Eng C 2012;32:127.
[29] Tehrani RMA, Salari AA. The study of dehumidifying of carbon monoxide and
ammonia adsorption by Iranian natural clinoptilolite zeolite. Appl Surf Sci
2005;252:86670.

564

A. Alshameri et al. / Journal of the Taiwan Institute of Chemical Engineers 45 (2014) 554564

[30] Vassileva P, Voikova D. Investigation on natural and pretreated Bulgarian


clinoptilolite for ammonium ions removal from aqueous solutions. J Hazard
Mater 2009;170:94853.
[31] Widiastuti N, Wu H, Ang HM, Zhang D. Removal of ammonium from greywater
using natural zeolite. Desalination 2011;277:1523.
[32] Townsend RP, Loizidou M. Ion exchange properties of natural clinoptilolite, ferrierite and mordenite: 1. Sodiumammonium equilibria. Zeolites 1984;4:1915.
[33] Lei L, Li X, Zhang X. Ammonium removal from aqueous solutions using microwave-treated natural Chinese zeolite. Sep Purif Technol 2008;58:35966.
[34] Wu Z, An Y, Wang Z, Yang S, Chen H, Zhu Z, et al. Study on zeolite enhanced
contactadsorption regenerationstabilization process for nitrogen removal. J
Hazard Mater 2008;156:31726.
[35] Gottardi G, Galli E. Natural zeolites. 1st ed. Berlin: Springer-Verlag; 1985.
[36] Demir A, Gunay A, Debik E. Ammonium removal from aqueous solution by
ion-exchange using packed bed natural zeolite. Water SA 2004;28:32936.
[37] Zaghouane-Boudiaf H, Boutahala M. Kinetic analysis of 2,4,5-trichlorophenol
adsorption onto acid-activated montmorillonite from aqueous solution. Int J
Miner Process 2011;100:728.
[38] Thomas JM, Thomas WJ. Principles and practice of heterogeneous catalysis.
3rd ed. Weinheim: Wiley-VCH; 1997.
[39] Moussavi G, Khosravi R. Removal of cyanide from wastewater by adsorption
onto pistachio hull wastes: parametric experiments, kinetics and equilibrium
analysis. J Hazard Mater 2010;183:72430.

[40] Ugurlu M, Karaoglu MH. Adsorption of ammonium from an aqueous solution


by y ash and sepiolite: isotherm, kinetic and thermodynamic analysis.
Microporous Mesoporous Mater 2011;139:1738.
[41] Nemr AE, Abdelwahab O, El-Sikaily A, Khaled A. Removal of direct blue-86
from aqueous solution by new activated carbon developed from orange peel. J
Hazard Mater 2009;161:10210.
[42] Liao P, Yuan S, Zhang W, Tong M, Wang K. Mechanistic aspects of nitrogenheterocyclic compound adsorption on bamboo charcoal. J Colloid Interface Sci
2012;382:7481.
[43] Shen XE, Shan XQ, Dong DM, Hua XY, Owens G. Kinetics and thermodynamics
of sorption of nitroaromatic compounds to as-grown and oxidized multiwalled carbon nanotubes. J Colloid Interface Sci 2009;330:18.
[44] Arami M, Limaee NY, Mahmoodi NM. Evaluation of the adsorption kinetics and
equilibrium for the potential removal of acid dyes using a biosorbent. Chem
Eng J 2008;139:210.
[45] Zhao Y, Zhang B, Zhang X, Wang J, Liu J, Chen R. Preparation of highly ordered
cubic NaA zeolite from halloysite mineral for adsorption of ammonium ions. J
Hazard Mater 2010;178:65864.
[46] Aksu Z, Kabasakal E. Batch adsorption of 2,4-dichlorophenoxy-acetic acid (2,4D) from aqueous solution by granular activated carbon. Sep Purif Technol
2004;35:22340.
[47] Manju GN, Raji C, Anirudhan TS. Evaluation of coconut husk carbon for the
removal of arsenic from water. Water Res 1998;32:306270.

Вам также может понравиться