You are on page 1of 8

Applied Clay Science 140 (2017) 157–164

Contents lists available at ScienceDirect

Applied Clay Science

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

Research paper

Adsorption of phosphate ions from aqueous solution by modified


bentonite with magnesium hydroxide Mg(OH)2
Mohamed El Bouraie a,⁎, Alaa A. Masoud b
a
Central Laboratory for Environmental Quality Monitoring (CLEQM), National Water Research Center (NWRC), El-Qanater El-Khairiya 13621, Egypt
b
Geology Department, Faculty of Science, Tanta University, Egypt

a r t i c l e i n f o a b s t r a c t

Article history: This study investigated phosphate ions removal from aqueous solutions by using modified bentonite with mag-
Received 3 August 2016 nesium hydroxide in batch system. Raw bentonite (RB) and Mg-modified bentonite (MB) were characterized by
Received in revised form 13 January 2017 Scanning Electron Microscopy (SEM), X-ray Diffraction (XRD), and Fourier Transform Infrared Spectroscopy (FT-
Accepted 19 January 2017
IR). Adsorption experiments were conducted on the adsorption of phosphate onto RB and MB in batch experi-
Available online 16 February 2017
ments. Phosphate ion removal by MB was pH dependent, and the optimum adsorption was observed at pH 7.
Keywords:
The adsorption process was relatively fast and equilibrium conditions were established within 120 min at
Adsorption 45 °C. The results were analyzed according to the Langmuir and Freundlich isotherm equations. The adsorption
Phosphate data is well interpreted by the Langmuir isotherm. Phosphate solution at a concentration of 25 mg/L was
Bentonite adsorbed by MB, and the final adsorption efficiency was N54%. The results showed that phosphate adsorption
Modification density of MB was high with the maximum adsorption density of 14.33 mg/m2, which suggested that MB was
Adsorption model an excellent adsorbent for effective phosphate removal from water. Thermodynamically negative ΔG°, positive
ΔH°, and positive ΔS° demonstrated the high affinity, and endothermic adsorption process between MB and
phosphate from aqueous solutions.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction removal methods such as chemical and biological treatments had


been successfully applied (Yeoman et al., 1988). Nevertheless, increas-
Contamination of surface water with phosphate and its removal in ing attention had been paid to adsorptive removal of phosphate from
water treatment had become increasing focus worldwide. Many aqueous solution (Hano et al., 1997; Donnert and Salecker, 1999). Eu-
dephosphorization studies had been made for aqueous solution, includ- trophication had become the primary water quality issue for most of
ing biological, chemical precipitation, and adsorption (Yu et al., 2010). the freshwater ecosystems in the world (Smith et al., 2006). N38% of
Of all phosphates removal techniques, adsorption was increasing atten- water bodies in many regions of the world were considered to have eu-
tion and becoming an attractive technology because of its simplicity, trophication problems. One gram of phosphorous was required for
low cost, ease of operation and handling, sludge free operation, and every 7 g of nitrogen for the formation of the organic matter created
the capacity of regenerate and reuse solids. In this regard, many adsor- in the process (Ross et al., 2008). Thus, the excess of bioavailable phos-
bents had been explored such as: zeolite; bauxite refinery residues (red phorous was the key nutrient which was understood to lead to eutro-
mud; BauxsolTM); calcined dolomite; fly ash; and ferric iron oxides phication of water bodies, resulting in increased aquatic plant and
(Fytianos et al., 1998; Wu et al., 2006; Huang et al., 2008; Despland et algal growth.
al., 2011). Chemical methods for removing phosphate compounds caused
Phosphorus (P) is an essential, often limiting, nutrient for growth of product quality reducing and side effects on aquatic environments.
organisms in most ecosystems. However, excessive supply of phospho- Among the available methods for removing these pollutants, adsorption
rus from wastewater into water bodies, such as lakes, rivers and creeks is still one of the most preferred methods, especially for effluents with
cause eutrophication, resulting in the bloom of aquatic plants, growth of moderate and low pollutant concentration (Nameni et al., 2007). In
algae and depletion of dissolved oxygen. Phosphate removal from aque- the past years, there had been increasing interest in developing recycla-
ous solution had been widely studied during the past decades. Typical ble inorganic adsorbents, particularly from bentonite for efficient re-
moval of organic pollutants from aqueous solutions (Nouri, 2002).
⁎ Corresponding author.
Raw bentonite is one of the abundant clay minerals at the earth's
E-mail addresses: mido.chemie@gmail.com (M. El Bouraie), surface, which can be used as an effective adsorbent for many toxic sub-
alaa_masoud@science.tanta.edu.eg (A.A. Masoud). stances in soil, water and air. Bentonite is effective adsorbent for cations

http://dx.doi.org/10.1016/j.clay.2017.01.021
0169-1317/© 2017 Elsevier B.V. All rights reserved.
158 M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164

but it shows lower affinity toward negative groups, like phosphate, due Table 1
to the absence of effective adsorption sites for anions in water. Adsorp- Physical and chemical characteristics of raw and modified bentonite.

tion onto bentonite that contains montmorillonite appears to involve Parameters Raw bentonite Modified bentonite
two distinct mechanisms: (i) an ion exchange reaction at permanent- Physical characteristics
charge sites, and (ii) formation of complexes with the surface hydroxyl Loss on ignition (%) 12.66 ± 0.06 c 12.31 ± 0.04 b
groups. Modified bentonite is prepared by exchanging the naturally oc- Density 1 ± 0.02 c 1.05 ± 0.02 a
curring interlayer monovalent and divalent cations like sodium and/or Granulometry (μm) 1–2 ± 0.03 c 1–1.5 ± 0.04 b
Capacity swelling Cg 8.31 ± 0.08 b 8.52 ± 0.06 c
calcium with highly charged polymeric metal species such as magne-
pH for 10 g/L 9.6 ± 0.02 c 7.8 ± 0.04 b
sium. Recently, raw bentonite was successfully employed for the ad- Conductance (μS) for 0.5 g/L 63.5 ± 0.03 c 81.4 ± 0.02 c
sorption of metal ions and dyes. It has been considered as a potential Specific surface area, BET (m2/g) 33.52 ± 0.02 b 51.72 ± 0.04 a
adsorbent for the removal of pollutants from water. The effective appli-
Chemical composition (wt%)
cation of bentonite for the water treatment is limited due to its surface SiO2 55.70 ± 0.03 c 59.5 ± 0.02 b
area and presence of net negative charge leading to its low adsorption Al2O3 21.5 ± 0.08 b 0.03 ± 0.05 c
capacity. As a result in the present research work modification of phys- Fe2O3 3.5 ± 0.09 a 3.65 ± 0.03 c
ical structure and chemical properties of bentonite results the maximi- MgO 5.6 ± 0.06 c 12.2 ± 0.05 a
TiO2 0.84 ± 0.08 b 0.92 ± 0.06 c
zation of its adsorption capacity (Tahir et al., 2010).
CaO 3.15 ± 0.04 c 3.26 ± 0.03 a
A composite adsorbent, modified bentonite with Magnesium Hy- PO4 0.75 ± 0.05 c 13.01 ± 0.06 b
droxide (MB), was proposed and studied in this research. The reason K2O 0.51 ± 0.04 c 0.36 ± 0.03 a
for choosing magnesium is that relative to Fe+3 or Al+3 ions, magne- Na2O 3.75 ± 0.08 b 2.89 ± 0.04 c
MnO 1.56 ± 0.02 c 1.08 ± 0.01 b
sium hydroxide has a higher affinity for phosphate (Fan and
FeO 0.82 ± 0.04 b 0.54 ± 0.05 b
Anderson, 2005). Bentonite, which had a high surface area, should pro- C 1.47 ± 0.02 c 1.62 ± 0.06 c
vide an efficient surface for magnesium hydroxide. At the same time, S 0.85 ± 0.03 a 0.94 ± 0.04 b
the efficiency of adsorption was greatly affected by the MB amount The same letters in the same column indicate no significant difference (p N 0.05).
and its preparation conditions. MB was known to have high surface
area and permanent porosity, which makes them attractive adsorbents
(Jiang et al., 2004). The use of MB in environmental applications was as on the outer surface and edges. The cation exchange capacity (CEC)
gained increasing attention due to the versatility of the modification of the sample was determined by the methylthioninium chloride meth-
process through which the porosity and properties of both pillar and od before purification was about 0.512 meq/g, this value was low com-
bearing material can be modified to achieve desired product properties pared to pure bentonite sample. Therefore, the sample was not pure
(Gurses et al., 2006). enough and the mass yield was 63.88% (Olu-Owolabiand and
The objective of this study was to examine the feasibility of using Unuabonah, 2011). The specific surface areas (BET) of the bentonite
Magnesium modified bentonite as adsorbents for phosphate removal samples were determined by ASAP 2000 micropore analysis. The BET
from aqueous solutions. A modified bentonite based on natural benton- of RB was 33.52 m2/g.
ite, was prepared and phosphate uptake was evaluated vs. pH using ad-
sorption efficiency and adsorption isotherms. The effects of temperature 2.3. Purified bentonite
and thermodynamic study on phosphate adsorption capability were
also investigated. RB was compared with MB used for phosphate ions Bentonite particles were dispersed in water and heated at 75 °C in
removal from aqueous solutions. the presence of a solution composed of 2 M AlCl3 and 1 M NaOH. The
purpose of this operation was to eliminate inorganic and organic com-
2. Materials and methods pounds, various free cations found in the interlayer spaces, and then sat-
urated with aluminium (Al3+) to ensure complete transformation into
2.1. Chemicals the aluminium form. After these exchanges, the slurry was stirred for
12 h at room temperature, filtered, and washed repeatedly with deion-
All chemicals used in the present study were of analytical reagent ized water (Arias and Sen, 2009). The suspension obtained was put in
grade. Sodium hydroxide and magnesium chloride were obtained dialysis membranes to remove the chloride ions adsorbed onto the sur-
from Merck, Germany. A stock solution of phosphate (40 mg/L) was face from the layers. The dialysis water was renewed until what the test
prepared by dissolving 175.75 mg of KH2PO4, dried at 120 °C for 2 h, with silver nitrate indicates the absence of chloride ions. Then, the puri-
in 1 L of deionized water (Conductivity = 0.5 μS/m). Working standards fied bentonite suspension was dried in an oven at a temperature not ex-
were prepared by dilution of the stock solution. 0.1 M HCl and 0.1 M ceeding 60 °C in order to be activated later. The CEC was determined by
NaOH solutions were used for pH adjustment. the methylene blue method after purification activation was about
1.13 meq/g and the mass yield was 55.62%.
2.2. Raw bentonite
2.4. Modification of bentonite by magnesium hydroxide
RB was used as an adsorbent for the experiments. All bentonite sam-
ples were in the clod sized, forms when first received. Latter, they were 10 g of treated bentonite was immersed in 150 mL of 2.0 M magne-
dried at 105 °C for 4 h in a drying air oven, and then were grounded sium hydroxide and the mixture was agitated at 90 °C for 6 h. The ob-
using a porcelain mill to 74 μm for the experiments. The results of the tained powder was rinsed with 0.01 M HCl aqueous solution to
mineralogical and chemical compositions of the materials are presented remove the excess Mg(OH)2 precipitated on the outer surface of ben-
in Table 1. tonite and further washed with deionized water. In some times, The col-
The basic structural unit of RB was composed of two tetrahedral co- our on surface bentonite was turned from original light colour to dark
ordinated sheets of silicon ions surrounding a sandwiched octahedral brown, indicating the oxidation of the hydroxide into oxide phase at
coordinated sheet of aluminium ions. The isomorphous substitution of room temperature and the yield was mashed manually (Liu et al.,
Al3+ for Si4+ in the tetrahedral sheet and Mg2+ for Al3+ in the octahe- 2011). Some inorganic materials can be used to improve the gelation
dral sheet results in a net negative surface charge on bentonite. Com- property of bentonite. According to this study, flowability and stability
pared with other clay types, it has excellent adsorption properties and of RB was improved by the addition of Mg(OH)2. However, it is worth
possesses adsorption sites available within its interlayer space as well to mention that the addition of Mg(OH)2 into the suspension
M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164 159

significantly increases the viscosity of bentonite suspensions. In addi- al., 2012). All experiments were repeatedly performed in duplicate.
tion, the solubility of Mg(OH)2 was too low, which led to regulate the The experimental error limit of duplicates was maintained at ±0.5%.
viscosity behaviours of bentonite suspensions. This was an interesting
point that the positive effect of Mg ions in bentonite suspensions, how- 2.6.1. Adsorption isotherm
ever, without emphasizing the effect of Mg(OH)2 particles in suspen- Langmuir and Freundlich isotherm models were used to establish
sions. Since Mg(OH)2 does not dissolve well in bentonite suspensions, the relationship between the amount of adsorbed phosphate ion onto
the effect of dissolved Mg ions on suspension viscosity can be eliminat- MB and its equilibrium concentration in aqueous system (Krishna and
ed. In this study, the modification process was performed to understand Bhattacharyya, 2002).
the interaction mechanism between bentonite particles and Mg(OH)2. Langmuir adsorption isotherm was based on the assumption of
Therefore, the BET surface area value increased from 33.52 to monolayer adsorption onto a surface with a finite number of identical
51.72 m2/g. The higher surface area was obtained because of the remov- sites (Sen and Sarzali, 2008). Langmuir isotherm could be arranged in
al of inorganic impurities by the Mg(OH)2 interaction. The higher values its linear form as Eq. (3):
of the specific surface area were related to the agglomerated structure of
magnesium hydroxide particles, which had been created during the Ce Ce 1
¼ þ ð3Þ
modification process. qe qm K L qm

2.5. Characterization of the adsorbents


Where Ce is the equilibrium concentration of phosphate (mg/L) and
qe is the amount of adsorbed in milligram per unit specific surface area
The morphological and chemical compositions of raw and modified
of adsorbent (mg/m2). qm and KL are Langmuir constant relating ad-
bentonite were obtained by the scanning electron microscope (SEM) a
sorption density (mg/m2) and the energy of adsorption (L/g), respec-
JSM6360LV SEM with High-Low vacuum (JEOL) with the accelerating
tively. These constants can be calculated from the slope and intercept
voltage of 20 kV with X-ray dispersive spectrometer (JXA8621
of the linear plots of Ce/qe versus Ce, respectively.
Superprobe; JEOL, Japan). The Fourier Transformed Infrared (FTIR)
The dimensionless parameter of adsorption (RL) (defined as RL =
spectra of the bentonites were obtained using KBr wafers and
1/(1+ KLCo), where Co is the initial concentration was used as an indi-
SHIMADZU 8400S FTIR. The FTIR was used to identify the surface func-
cator to assess the extent of adsorption). Depending on the RL value,
tional groups of two bentonites.
there are four possibilities for adsorption: (1) favorable adsorption if
0 b RL b 1, (2) unfavorable adsorption when RL N 1, (3) linear adsorption
2.6. Phosphate adsorption experiments
for RL = 1, and (4) irreversible adsorption for RL = 0 (Sen and Dustin,
2011).
The adsorption of PO3−
4 ions onto adsorbents were tested in a batch
The adsorption data were also fitted to Freundlich isotherm, which is
equilibration by mixing 2 g of the sorbent in a 1 L phosphate solution
described by the linear form following Eq. (4):
with the initial concentrations (ranging from 0.05 to 25 mg/L). The ex-
periments were carried out at pH of 6.0 and 7.0 for both bentonites  
1
(RB and MB) under the operating temperature (45 ± 0.1 °C), and agita- logqe ¼ logK F þ logC e ð4Þ
tion speed of 200 rpm within the equilibrium time of 120 min, unless n
otherwise specified by the study design (i.e., to examine the effects of
pH, temperature, initial metal concentration, and contact time). After Where KF and n are Freundlich constants incorporating all factors af-
equilibration, supernatant were withdrawn and filtered with 0.45 μm fecting the adsorption density and intensity of adsorption, respectively.
syringe driven filter (Millex-LH, PTFE, Millipore Corp., Ireland). The fil- Values of KF and n were determined from the intercept and slope of the
trate concentration was measured by the molybdate blue spectrophoto- linear plot of log qe versus log Ce.
metric method (APHA, 2005) using a Lambda 25 UV/VIS spectro-
photometer (Perkin-Elmer, Germany). The determination limit of the 2.6.2. Thermodynamic parameters
analytical method was 0.01 mg PO−3 4 /L. Blank samples with no adsor- The removal of phosphate ions by MB was studied from thermody-
bent were perpetrated and monitored as a control. All experiments namic viewpoint to ascertain the nature of adsorption process under
were run in triplicate. Phosphate removal efficiency (E) was calculated the condition of the current study. To achieve this goal, three thermody-
according to Eq. (1): namic parameters, including Gibb's free energy (ΔG°), enthalpy change
(Δ H°) and change in entropy (Δ S°), were determined by using Eqs.
ðC o −C e Þ (5)–(8):
E¼  100 ð1Þ
Co
Cs
KL ¼ ð5Þ
Ce
Where E is the phosphate removal efficiency (%), Co is the initial con-
centration of phosphate (mg PO− 3
4 /L) and Ce is the equilibrium
concentration. ΔG ° ¼ −RT ln K L ð6Þ
The phosphate adsorption kinetics were studied after specified peri-
od of time the flasks were taken out for the analysis of phosphate con- ΔS ° ΔH °
ln K L ¼ − ð7Þ
centration. The amount of adsorbed in milligram per unit specific R RT
surface area of adsorbent is called by adsorption density, qe
(mg PO−3 2
4 /m ) was commonly represented by using Eq. (2): ΔG ° ¼ ΔH °−TΔS ° ð8Þ

V ðC o −C e Þ Where Cs and Ce are the removed and remaining concentrations, re-


qe ¼ ð2Þ
BET spectively. KL is the distribution coefficient for the adsorption, ΔS°, ΔH°,
and ΔG° are the changes of entropy, enthalpy, and the Gibbs energy, T
Where V is the volume of solution (L), Co and Ce are the concentra- (K) is the temperature, and R (8.314 J/mol K) is the gas constant. The
tion of phosphate (mg PO−34 /L) before and after adsorption, respective- values of ΔH° and ΔS° were determined from the slopes and intercepts
ly, and BET (m2/g) is the specific surface areas of adsorbent (Zehhaf et of the plots of ln KL versus 1/T (Eloussaief et al., 2011; Özdes et al., 2011).
160 M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164

of oxy-silica-montmorillonite was reduced and (d200) reflection of dolo-


mite disappeared. However, the shift of (d001) reflection of MB to the
right on x axis points to the phosphate ion exchange from interlayer
space for anions.
Increasing ratio of Mg(OH)2/bentonite also increased the number of
active sites MgO which has high catalytic activity and basicity. The ac-
tive site MgO is responsible for the presence of the magnesium hydrox-
Fig. 1. SEM micrographs of bentonites: raw bentonite (a), Mg-modified bentonite (b) and
ide molecule between the silicate layers. However, with further increase
bentonite after adsorption (c).
of Mg(OH)2 loading, the interactions between Mg(OH)2 with internal
layer of bentonite (Si\\O\\Al stretching groups) were excessive and
3. Results and discussion during the modification, a new phase of Si\\O\\Mg compound was
formed. This new phase compound (Si\\O\\Mg) had lower catalytic ac-
3.1. Characterization of adsorbents tivity and basicity than the MgO phase (Rahni et al., 2014). The XRD pat-
tern of MB showed poor crystallinity, broad and less intense peaks as
SEM images help understand the microscale surface morphology of compared to the parent bentonite mineral due to the presence excess
the bentonite samples. As can be seen in Fig. 1, raw bentonite surface amount of Si-O-Mg compound, or an irregular stacking of the activated
is relatively more smooth and reveal spongy appearance with irregular and non-activated layers and thus the structure of the resultant benton-
structure than modified bentonite after the treatment with magnesium ite became amorphous (Eren and Afsin, 2008). From XRD diagrams, the
ions, which lead to the formation of Mg(OH)2 clusters between the in- peaks present as in Fig. 2c, attributed to the physical adsorbed PO3−
4 ions
terlayer spaces of bentonite. These Mg(OH)2 clusters formed between onto MB. According to the intensity and quantity of the diffraction
the interlayers are rigid enough not only to prevent the interlayer spac- peaks, it can be speculated that Mg(OH)2 on the surface of MB was grad-
ing from collapsing, but also to generate pores larger than those of con- ually converted in situ into Mg3(PO4)2. In order to clarify the mecha-
ventional zeolites. After adsorption, modified bentonite surfaces nism of adsorption of PO3− 4 ions onto MB as shown in (Scheme 1).
became swollen. This swelling could be the result of phosphate The FTIR spectra of raw and modified bentonite samples in the
adsorption. wavenumber range of 4000–500 cm−1 were shown in Fig. 3. The spec-
The X-ray diffraction analysis (XRD) was mainly used to give data trum of raw bentonite exhibited absorption bands at 3450 and
pertinent to basal spacing of minerals and established the fact of adsorp- 1650 cm−1 this was assigned to the stretching and bending vibrations
tion and provided information regarding the absorbed molecules into of the OH groups for the water molecules adsorbed on bentonite sur-
the basal spacing of bentonite particles. XRD analysis was carried out face, and a band at 3621 cm−1 which represented the stretching vibra-
by mixing the modified and raw bentonite separately with magnesium tion of the hydroxyl groups coordinated to octahedral Al3 + cations
hydroxide to evaluate the changes occurred in the interlaminar distance (Sdiri et al., 2011). The raw bentonite spectrum also contained a band
of bentonite resulting from their inter action with Mg(OH)2. The X-ray between 693 and 796 cm−1 which were attributed to orthodase and
diffractograms of raw, Mg-modified bentonite and bentonite after ad- quartz respectively (Sdiri et al., 2010). The intensive band at
sorption of phosphate ions were shown in Fig. 2. The small angle (3°– 1088 cm−1 was assigned to the Si\\O stretching vibration, whereas
10°) of XRD diffraction peaks could be used to study the structure of the bands around 466 and 521 cm− 1 were ascribed to Si\\O\\Al
mesoporous materials. XRD results showed that the exception of SiO2 (where Al was the octahedral cation) and Si\\O\\Si bending vibrations,
after modification of the adsorbent, other compounds of the adsorbent respectively. The intensive bands near 521, 850 and 916 cm−1 attribut-
decreased (Flessner et al., 2001). Until a certain loading, Al2O3 were con- ed to Al\\Mg\\OH, Al\\Al\\OH and Al\\OH\\O\\Si stretching vibra-
verted to AlCl3 during the modification process. Fig. 2a shows that peaks tion respectively. FTIR spectra of Fig. 3b and c showed the decreasing
observed around 2θ = 22°, 24°, 27°, 31°, 36° and 42° are belong to Al2O3 intensity as a result of modification bentonite with magnesium hydrox-
phase. Bentonite samples were mainly composed of dioctahedral (SiO2) ide which reflects the leaching of octahedral cation, such as Al3+ from
and specifically montmorillonite with the d001 basal reflection exhibited the bentonite structure. The vibrational band near 3626 cm−1, assigned
at about 15.3 Å (Fig. 2a). The estimated montmorillonite contents of the to stretching Mg-OH vibrations is typical for Mg-rich dioctahedral lay-
samples were found 91%, and 86% for the samples RB and MB, respec- ered minerals. The reduction of the content of the octahedral cation
tively. Thus, both samples were characterized by high concentration of was accompanied by a decrease of both OH bending vibrations at 850
montmorillonite and low levels of impurities. The main (d001) reflection and 916 cm−1. The stretch vibrations found within band range 1100–

Fig. 2. X-ray diffraction patterns of bentonites: raw bentonite (a), Mg-modified bentonite (b) and bentonite after adsorption (c).
M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164 161

Scheme 1. Adsorption mechanism of phosphate onto modified bentonite.

1200 cm− 1 appeared to represent functional group of phosphorus


(P_O) belonging to phosphate ions as in Fig. 3c. The FTIR result was
in clear agreement with the SEM and XRD studies which indicate se-
quential degradation of the bentonite sheet upon acid treatment.
Fig. 4. Effect of pH on phosphate-uptake by raw and modified bentonites.

contradiction to the expected; the latter was attributed to the electro-


3.2. Effect of pH
static attraction mechanism (solution pH b pH during the initial adsorp-
tion phase). PZC could change when the surface functional groups are
Adsorption of phosphate by raw and modified bentonite was inves-
altered due to their modification. The surface charge was effected by
tigated in the pH range 4–9 while maintaining the other parameters
the solution pH; at pH b pHPZC the surface was positively charged,
constant. Fig. 4 showed the change in phosphate uptake by both ben-
while at pH N pHPZC was negatively charged. Hence, the chemical mod-
tonites at different initial pH levels. In general, Raw bentonite carry
ification of bentonite with Mg(OH)2 and the enhanced chemical or/and
two types of electrical charge; one permanent negative charge generat-
physical adsorption mechanisms should play a greater role on removal
ed from its structural components and a second variable surface charge
of phosphate (Mekhloufi et al., 2013).
which depends on the pH of the liquid phase.
The pH-dependent increase was due to increasing adsorption of
The adsorption efficiency of MB increased gradually in the pH range
formed phosphate anions, with pKa 7.2 (Papadas et al., 2009), and
of 4–7 and reached a maximum value (54.6%) when the pH value was 7.
with the positively charged surface sites of bentonite. Meanwhile, the
At pH values over 8, the surface of MB became negatively charged thus
major decrease in adsorption was due to the repulsion between phos-
phosphate adsorption drops dramatically (13.4%). Although the in-
phate molecules and the abundance of OH− ions at higher pH values
creased adsorption efficiency in this region was very pronounced for
(Stathi et al., 2007). Another reason for this decrease was due to the par-
the MB, such increment was not observed for RB. The adsorption effi-
ticle interactions resulting from the high adsorbents dose, such as ag-
ciency of RB changed very little within the pH range of 4–6 and de-
gregations. Aggregations could lead to a decrease in the total surface
creased in solutions with higher pH values. Hence, with increased of
area of the adsorbents and an increase in the diffusional path length.
pH values, the negative charges of the RB surface that made electrostatic
However, phosphate adsorption onto MB appeared to be more pH de-
repulsion between phosphate ions with SiO2 which had been formed at
pendent than with RB.
high pH. But in acidic pH, the surface is positively charged due to the
high hydrated radius, that made electrostatic attraction with PO−34 ions.
In the present study, the Mg(OH)2 modified of bentonite resulted to 3.3. Effect of contact time
an increase of pH value at point of zero charge (PZC) from 8.4 to 9.7; that
means for a wide range of pH values, the MB remains positively charged, Fig. 5 showed the change in phosphate adsorption efficiency with
favoring the electrostatic attraction to H2PO− 4 (pKa = 2.1) and to time on raw and modified bentonite. Adsorption studies were contin-
HPO−24 (pKa = 7.2) species. As we have already shown, the initial PZC ued for 250 min, and it was found that adsorption efficiency increased
of the RB was accompanied with a low affinity for phosphate in with the increasing contact time, and a large portion of the phosphate
ions were adsorbed onto RB within a few minutes (20 min); whereas
MB reached equilibrium almost at the end of 120 min. This was due to
MB had more expanded sheets between the layers than RB, confirmed
the significant role of chemical adsorption as a removal mechanism
for phosphate. In MB, phosphate binding was lower at the beginning
of the experiment, probably because of the surface interaction (physical

Fig. 3. FTIR spectra of bentonites: raw bentonite (a), Mg-modified bentonite (b) and
bentonite after adsorption (c). Fig. 5. Effect of the contact time on phosphate-uptake by raw and modified bentonites.
162 M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164

Fig. 6. Effect of phosphate concentration on adsorption density of raw and modified Fig. 8. Effect of temperature on adsorption densities of raw and modified bentonites
bentonites. toward phosphate.

adsorption) combined with slow dissolution of Mg(OH)2 that usually of PO−3


4 ion increased with the increase in the adsorbent amount (ad-
constitutes the first step of phosphate adsorption. The adsorption may sorption density) due to instauration of the active sites on the adsorbent
take place on the outer surface and in the interlayers of the bentonites. surface during the adsorption process (Zamparas et al., 2012).
Therefore, phosphate uptake at the beginning was about 2.3 times
higher for MB than RB. However, adsorption ability for MB reached 3.6. Effect of temperature
five times higher value than that of RB at the end of the experiments.
The effect of temperature on phosphate adsorption was determined
at temperatures ranging from 25 to 50 °C. As shown in Fig. 8, high tem-
3.4. Effect of phosphate concentration
perature was advantageous for phosphate adsorption on both benton-
ites. Fig. 8 showed an increase in the adsorption density with an
The removal of phosphate ions onto the adsorbents as a function of
increase in temperature. This indicated that the adsorption reaction
their concentrations was investigated at constant temperature (45 ±
was of endothermic nature and the ion-exchange mechanism was fa-
0.1 °C) by varying the initial phosphate concentration from 0.05 to
vored at higher temperatures. The increases in adsorption density of
25 mg/L while keeping all other parameters constant. Fig. 6 represented
MB at higher temperatures could be caused by the enlargement of
the change in adsorption behavior of both bentonites under different
pore size and/or modification of the adsorbent surface (Yan et al.,
phosphate ion concentrations. The adsorption density increased with
2010). Likewise, decreased adsorption density of both bentonites at
the increasing amounts of phosphate added onto RB and MB. At a low
lower temperatures, could be caused by decreased pore size. Conse-
rate of phosphate added (0.05 mg/L), the amounts of PO−3 4 adsorbed
quently, the maximum amounts of phosphate adsorbed on RB and MB
onto RB was 0.002 mg/m2 (3.0%). However, at the high phosphate
were increased from 0.72 to 1.16 mg/m2 and 10.33 to 14.02 mg/m2, re-
added (25 mg/L) rate, the amounts of PO−3 4 adsorbed onto RB was
spectively. It was common that increasing temperature could create a
1.62 mg/m2 (6.49%). But in case of MB, amounts of phosphate adsorbed
swelling influence inside the adsorbent structure.
at the low (0.05 mg/L) was 0.015 mg/m2 (30%) and high (25 mg/L) rate
added was 14.33 mg/m2 (57.32%). Increasing initial phosphate concen-
3.7. Adsorption isotherm
tration led to increase the PO− 4
3
uptake onto adsorbents to a certain
point; then, a plateau occurred for both bentonites that indicated un-
Adsorption isotherms described the distribution of phosphate ion on
availability of adsorption sites (Kaya and Ören, 2005).
RB and MB at equilibrium. Adsorption balance of phosphate ions was
studied by the Freundlich and Langmuir adsorption isotherm. Langmuir
3.5. Effect of adsorbent amount isotherm was the indicator of active surface adsorption on the homoge-
nous surface, while the Freundlich isotherm was used for heteroge-
The effect of adsorbent amount on the sorption of phosphate was neous surfaces (Dada et al., 2012). The linear form of Langmuir and
analyzed kinetically over a range of 0.2 to 10 g. The percentages of Freundlich equations were as below.
adsorbed phosphate increased from 0.57% to 5.44% onto RB and from Fig. 9 illustrated the plots of phosphate ion where a straight line
7.46% to 54.42% onto MB, respectively (Fig. 7). The removal efficiency could be well observed between Ce/qe and Ce. This indicated that the ex-
perimental data followed Langmuir's isotherm. qm and KL were

Fig. 9. Langmuir isotherm plots for the adsorption of phosphate onto raw and modified
Fig. 7. Effect of adsorbent amount on phosphate-uptake by raw and modified bentonites. bentonites.
M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164 163

Table 2
Langmuir and Freundlich isotherm parameters for phosphate ions adsorption onto Raw
Clay and Modified Bentonite.

Adsorbent Temperature Langmuir isotherm Freundlich isotherm


(°C)
qm KL R2 KF n R2
(mg/m2) (L/mg)

25 0.391 0.076 0.391 0.034 1.019 0.836


Raw Ben 35 0.408 0.131 0.669 0.042 1.083 0.842
45 0.445 0.145 0.719 0.044 1.117 0.843
25 2.977 0.170 0.922 0.336 1.18 0.913
Mod Ben 35 3.137 0.172 0.937 0.37 1.193 0.927
45 3.435 0.223 0.967 0.38 1.264 0.961

determined from slope and intercepts of the plots (Fig. 9) and were pre-
sented in Table 2. From the results, it was clear that the value of adsorp-
Fig. 11. Van't Hoff plot of adsorption equilibrium constant KL.
tion efficiency qm and adsorption energy b of the adsorbent increased
with temperature. High temperatures increased the kinetic energy of
phosphate and, hence, enhanced the mobility of phosphate ion. This adsorption of phosphate through the percolation process could not be
led to a higher chance of phosphate being adsorbed onto the adsorbent ruled out. However, the values of n were greater than one indicating
and an increase in its adsorption density. Further, it confirmed the endo- the adsorption was much more favorable (Ho et al., 2002).
thermic nature of the processes involved in the system. To confirm the As both Langmuir and Freundlich isotherm models could explain the
adorability of the adsorption process, the separation factor (RL) had cal- adsorption, this suggested that the adsorption of phosphate was poten-
culated. The values were found between 0 and 1 and confirm that the tially monolayer (Feng et al., 2009). The values of 1/n b 1 confirmed a fa-
ongoing adsorption process was favorable (Sen and Dustin, 2011). vorable adsorption onto microporous adsorbent (Gunay et al., 2007).
A further analysis of the Langmuir equation could be made on the The results indicated that MB had higher maximum adsorption density
basis of a dimensionless equilibrium parameter, RL, also known as the for phosphate. Moreover, adsorption density for MB was found to have a
separation factor, given by (Sen and Khoo, 2013). higher KF value than RB, indicated that this phosphate had higher affin-
The separation factor, RL had been calculated from Langmuir plot. It ity for MB than RB. Overall Langmuir isotherm model had higher regres-
had been found that the calculated range of RL values from 0.133 to sion coefficient (R2) compared to the Freundlich isotherm model for
0.478 for RB and 0.158 to 0.529 for MB system with the initial phosphate both the systems.
ion range of 0.05 to 25 mg/L. These RL values indicated favorable adsorp-
tion as it lie in the range 0 b RL b 1. The maximum Langmuir adsorption
density of MB was more than RB (Sen and Khoo, 2013). 3.8. Adsorption thermodynamics
The Freundlich isotherm model was considered to be appropriate for
describing both multilayer sorption and sorption on heterogeneous sur- The thermodynamic equilibrium parameters (ΔG°, ΔH° and ΔS°) for
faces (Dada et al., 2012). The Freundlich isotherm was linear if 1/n = 1 the adsorption phosphate onto both bentonites at 45 °C were calculated
and, as 1/n decreased, the isotherm became more nonlinear. The units at various temperatures from the fit of the adsorption isotherms. A plot
for qe and Ce should be consistent if parameters were to have any prac- of ln KL versus 1/T was found to be linear as shown in Fig. 11. ΔH° and
tical application (Dada et al., 2012). Linear plot of log qe versus log Ce ΔS° were determined from the slops and intercepts of the plot (Fig.
showed that the adsorption of phosphate ion followed the Freundlich 11). For raw bentonite, the negative value of ΔG° reflects a more ener-
isotherm (Fig. 10). Values of KF and 1/n were found in the Table 2, getically favorable adsorption at low temperature (25 °C) as presented
showed the increase of positive charge on the surface that enhanced in Table 3. The decrease in Δ G° with the increase in temperature
the electrostatic force between MB surface and phosphate ion, which in- showed an increase in feasibility of sorption at higher temperature
creases the adsorption of phosphate. and so, indicated the nature of phosphate adsorption was not
The values clearly showed that dominance in adsorption density. spontaneous.
The intensity of adsorption was an indicative of the bond energies be- On the other hand, the negative value of ΔG° indicated the spontane-
tween phosphate and adsorbent and the possibility of slight chemical ous nature of phosphate adsorption ions on MB. The positive value of
sorption rather than physical sorption. The possibility of multilayer ΔH° indicated that the process of adsorption was endothermic and irre-
versible, probably due to nonpolar interactions. The positive value of
ΔS° suggested a high degree of disorderliness at the solid-solution inter-
face during the adsorption process of phosphate onto MB. It also
reflected the affinity of the adsorbent for phosphate and suggested
some structural changes in adsorbate and adsorbent (Donat et al.,
2005).

Table 3
Thermodynamics parameters of phosphate ions adsorption onto Raw Clay and Modified
Bentonite.

Adsorbent Temperature (°C) ΔG° (kJ/mol) ΔH° (kJ/mol) ΔS° (J/mol K)

25 −4.236 −8.246 −13.457


Raw Ben 35 −4.101 −8.246 −13.457
45 −3.966 −8.246 −13.457
25 −3.737 25.189 97.066
Mod Ben 35 −4.707 25.189 97.066
Fig. 10. Freundlich isotherm plots for the adsorption of phosphate onto raw and modified
45 −5.678 25.189 97.066
bentonites.
164 M. El Bouraie, A.A. Masoud Applied Clay Science 140 (2017) 157–164

4. Conclusions Gurses, C., Dogan, M., Yalcin, M., 2006. Adsorption kinetics of the cationic dye methylene
blue onto clay. J. Hazard. Mater. 131, 217–228.
Hano, T., Takanashi, H., Hirata, M., Urano, K., Eto, S., 1997. Removal of phosphorus from
Modified bentonite with Mg(OH)2 was an adsorbent for phosphate wastewater by activated alumina adsorbent. Water Sci. Technol. 35, 39–46.
ions removal from aqueous solutions. RB and MB were used to remove Ho, Y.S., Porter, J.F., McKay, G., 2002. Equilibrium isotherm studies for the sorption of di-
valent metal ions onto peat: copper, nickel and lead single component systems.
phosphate from aqueous solution. The phosphate adsorption capabili- Water Air Soil Pollut. 141, 1–33.
ties of RB were low, but high for MB. The pH effect, adsorbents concen- Huang, W., Wang, S., Zhu, Z., Li, L., Yao, X., Rudolph, V., Haghseresht, F., 2008. Phosphate
tration, equilibrium time, adsorption isotherms and thermodynamic removal from wastewater using red mud. J. Hazard. Mater. 158, 35–42.
Jiang, J.Q., Zeng, Z., Pearce, P., 2004. Evaluation of modified clay coagulant for sewage
adsorption, were examined. The results indicated that the modification
treatment. Chemosphere 56, 181–185.
process has significant effect on the removal efficiency of phosphate. Kaya, A., Ören, A.H., 2005. Adsorption of zinc from aqueous solutions to bentonite.
Moreover, the highest adsorption density was found with 25 mg/L ini- J. Hazard. Mater. B125, 183–189.
Krishna, D.G., Bhattacharyya, G., 2002. Adsorption of methylene blue on kaolinite. Appl.
tial phosphate ion concentration, at 45 °C and pH 7, as 14.33 mg/m2. Ad-
Clay Sci. 20, 295–300.
sorption equilibrium findings fitted with the Langmuir isotherm better Liu, Y., Li, Q., Gao, S., Shang, J.K., 2011. Exceptional As(III) sorption capacity by highly po-
when compared to the Freundlich isotherm. Thermodynamically nega- rous magnesium oxide nanoflakes made from hydrothermal synthesis. J. Am. Ceram.
tive ΔG°, positive ΔH°, and positive ΔS° demonstrated the high affinity Soc. 94, 217–223.
Mekhloufi, M., Zehhaf, A., Benyoucef, A., Quijada, C., Morallon, E., 2013. Removal of 8-
and endothermic adsorption process between the adsorbent and the quinolinecarboxylic acid pesticide from aqueous solution by adsorption on activated
adsorbed. The adsorbent used in this study demonstrated a relatively montmorillonites. Environ. Monit. Assess. 185, 10365–10375.
good phosphate adsorption density and removal yield when compared Nameni, M., Alavi Moghaddam, M.R., Arami, M., 2007. Adsorption of hexavalent chromi-
um from aqueous solutions by wheat bran. Int. J. Environ. Sci. Technol. 5, 161–168.
to studies conducted with bentonite modified by different methods or Nouri, S., 2002. Adsorption of aromatic compounds by activated carbon: effects of func-
with other adsorbents. However, RB was less efficient compared to tional groups and molecular size. Adsorpt. Sci. Technol. 20, 1–14.
MB, so that MB was a proper alternative for the removal of contami- Olu-Owolabiand, B.I., Unuabonah, E.I., 2011. Adsorption of Zn2+ and Cu2+ onto sulphate
and phosphate-modified bentonite. Appl. Clay Sci. 51, 170–173.
nants in processes that need large quantities of adsorbents, because of Özdes, D., Duran, C., Senturk, H.B., 2011. Adsorptive removal of Cd(II) and Pb(II) ions from
its suitable characteristics, such as availability, inexpensive, reusability aqueous solutions by using Turkish illitic clay. J. Environ. Manag. 92, 3082–3090.
and proper ability in removing contaminants, its efficiency would be in- Papadas, I.T., Kosma, C., Deligiannakis, Y., 2009. Ternary [Al2O3–electrolyte–Cu] species:
EPR spectroscopy and surface complexation modeling. J. Colloid Interface Sci. 339,
creased more, due to increase in the surface area. 19–30.
Rahni, S.Y., Mirghaffari, N., Rezaei, B., Ghaziaskar, H.S., 2014. Removal of phosphate from
Acknowledgement aqueous solutions using a new modified bentonite-derived hydrogel. Water Air Soil
Pollut. 225, 1–12.
Ross, G., Haghseresht, F., Cloete, T.E., 2008. The effect of pH and anoxia on the perfor-
Both authors would like to thank the staff of Central Laboratory for mance of phoslock, a phosphorus binding clay. Harmful Algae 7, 545–555.
Environmental Quality Monitoring for their cooperation during mea- Sdiri, A., Higashi, T., Hatta, T., Jamoussi, F., Tase, N., 2010. Mineralogical and spectroscopic
surements for providing necessary facilities to accomplish the work. characterization, and potential environmental use of limestone from the Abiod for-
mation, Tunisia. Environ. Earth. Sci. 61, 1275–1287.
Special thanks devoted to Central Laboratory for Tanta University to Sdiri, A., Higashi, T., Hatta, T., Jamousssi, F., Tase, N., 2011. Evaluating the adsorptive capac-
support present research and the required devices. ity of montmorillonitic and calcareous clays on the removal of several heavy metals
in aqueous systems. Chem. Eng. J. 172, 37–46.
Sen, T.K., Dustin, G., 2011. Adsorption of zinc (Zn+2) from aqueous solution on natural
References bentonite. Desalination 267, 286–294.
Sen, T.K., Khoo, C., 2013. Adsorption characteristics of zinc (Zn2+) from aqueous solution
American Public Health Association (APHA), 2005. Standard Method for Examination of by natural bentonite and kaolin clay minerals: a comparative study. Computat.
Water and Wastewater. 21st ed. (Washington DC, USA). Water, Energ, and Environ. Eng. 2, 1–6.
Arias, F., Sen, T.K., 2009. Removal of zinc metal ion (Zn+2) from its aqueous solution by Sen, T.K., Sarzali, M.V., 2008. Removal of cadmium metal ion (Cd+2) from its aqueous so-
kaolin clay mineral: a kinetic and equilibrium study. Colloids Surf. A Physicochem. lution by aluminium oxide: a kinetic and equilibrium study. Chem. Eng. J. 142,
Eng. Asp. 348, 100–108. 256–262.
Dada, A.O., Olalekan, A.P., Olatunya, A.M., 2012. Langmuir, Freundlich, Temkin and Smith, V.H., Joye, S.B., Howarth, R.W., 2006. Eutrophication of freshwater and marine eco-
Dubinin–Radushkevich isotherms studies of equilibrium sorption of Zn unto phos- systems. Limnol. Oceanogr. 51, 351–355.
phoric acid modified rice husk. J. Appl. Chem. 3, 38–45. Stathi, P., Litina, K., Gournis, D., Giannopoulos, T.S., Deligiannakis, Y., 2007. Physicochem-
Despland, L.M., Clark, M.W., Vancoc, T., Erler, D., Aragno, M., 2011. Nutrient and trace- ical study of novel organoclays as heavy metal ion adsorbents for environmental re-
metal removal by Bauxsol pellets in wastewater treatment. Environ. Sci. Technol. mediation. J. Colloid Interface Sci. 316, 298–309.
45, 5746–5753. Tahir, H., Uzma, H., Sultan, M., Jahanzeb, Q., 2010. Batch adsorption technique for the re-
Donat, R., Akdogan, A., Erdem, E., Cetisli, H., 2005. Thermodynamics of Pb2+ and Ni2+ ad- moval of malachite green and fast green dyes by using montmorillonite clay as adsor-
sorption onto natural bentonite from aqueous solutions. J. Colloid Interface Sci. 286, bent. Afr. J. Biotechnol. 9, 8206–8214.
43–52. Wu, D., Zhang, B., Li, C., Zhang, Z., Kong, H., 2006. Simultaneous removal of ammonium
Donnert, D., Salecker, M., 1999. Elimination of phosphorus from municipal and industrial and phosphate by zeolite synthesized from fly ash as influenced by salt treatment.
waste water. Water Sci. Technol. 40, 195–202. J. Colloid Interface Sci. 304, 300–306.
Eloussaief, M., Kallel, N., Yaacoubi, A., Benzina, M., 2011. Mineralogical identification, Yan, L.G., Xu, Y.Y., Yu, H.Q., Xin, X.D., Wei, Q., Du, B., 2010. Adsorption of phosphate from
spectroscopic characterization, and potential environmental use of natural clay mate- aqueous solution by hydroxy-aluminum, hydroxy-iron and hydroxy-ironaluminum
rials on chromate removal from aqueous solutions. Chem. Eng. J. 168, 1024–1031. pillared bentonites. J. Hazard. Mater. 179, 244–250.
Eren, E., Afsin, B., 2008. An investigation of Cu (II) adsorption by raw and acid-activated Yeoman, S., Stephanson, T., Lester, J.N., 1988. The removal of phosphorus during wastewa-
bentonite: a combined potentiometric, thermodynamic, XRD, IR, DTA study. ter treatment: a review. Environ. Pollut. 49, 183–233.
J. Hazard. Mater. 151, 682–691. Yu, Y., Wu, R.P., Clark, M., 2010. Phosphate removal by hydrothermally modified fumed
Fan, H.J., Anderson, P.R., 2005. Copper and cadmium removal by Mn oxide-coated granu- silica and pulverized oyster shell. J. Colloid Interface Sci. 350, 538–543.
lar activated carbon. Sep. Purif. Technol. 45, 61–67. Zamparas, M., Gianni, A., Stathi, P., Deligiannakis, Y., Zacharias, I., 2012. Removal of phos-
Feng, N.C., Guo, X.Y., Liang, S., 2009. Adsorption study of copper(II) by chemically modi- phate from natural waters using innovative modified bentonites. Appl. Clay Sci. 62–
fied orange peel. J. Hazard. Mater. 164, 1286–1292. 63, 101–106.
Flessner, U., Jones, D.J., Rozière, J., Zajac, J., 2001. A study of the surface acidity of acid- Zehhaf, A., Quijada, C., Benyoucef, A., Taleb, S., Berenguer, R., Morallon, E., 2012. Lead ion
treated montmorillonite clay catalysts. J. Mol. Catal. A Chem. 168, 247–256. adsorption from aqueous solutions in modified Algerian montmorillonites. J. Therm.
Fytianos, K., Voudrias, E., Raikos, N., 1998. Modelling of phosphorus removal from aque- Anal. Calorim. 110, 1069–1077.
ous and wastewater samples using ferric iron. Environ. Pollut. 101, 123–130.
Gunay, A., Arslankaya, E., Tosun, I., 2007. Lead removal from aqueous solution by natural
and pretreated clinoptilolite: adsorption equilibrium and kinetics. J. Hazard. Mater.
146, 362–371.

You might also like